Computational Investigation of Chemical Reactions ...

129
Computational Investigation of Chemical Reactions: Exploring the Reactivity of Nickel Enoate and Fumarate Complexes and Radical Assistance in the Force-Enabled Bond Scission of Poly(Acrylic Acid) by Michael T. Robo A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy (Chemistry) in the University of Michigan 2020 Doctoral Committee: Professor John Montgomery, Co-Chair Associate Professor Paul M. Zimmerman, Co-Chair Professor Brian J. Love Professor Melanie S. Sanford Professor John P. Wolfe

Transcript of Computational Investigation of Chemical Reactions ...

Page 1: Computational Investigation of Chemical Reactions ...

Computational Investigation of Chemical Reactions: Exploring the

Reactivity of Nickel Enoate and Fumarate Complexes and Radical

Assistance in the Force-Enabled Bond Scission of Poly(Acrylic Acid)

by

Michael T. Robo

A dissertation submitted in partial fulfillment

of the requirements for the degree of

Doctor of Philosophy

(Chemistry)

in the University of Michigan

2020

Doctoral Committee:

Professor John Montgomery, Co-Chair

Associate Professor Paul M. Zimmerman, Co-Chair

Professor Brian J. Love

Professor Melanie S. Sanford

Professor John P. Wolfe

Page 2: Computational Investigation of Chemical Reactions ...

Michael T. Robo

[email protected]

ORCID iD: 0000-0003-1356-246X

© Michael T. Robo 2020

Page 3: Computational Investigation of Chemical Reactions ...

ii

Dedication

To the brave and tireless healthcare workers involved in the ongoing COVID-19 pandemic.

Page 4: Computational Investigation of Chemical Reactions ...

iii

Acknowledgements

First, I would like to thank my parents, Jim Robo and Meredith Trim. They have provided me with

unending love and support for the entirety of my existence. My father pushed me early on to intern

at The Scripps Research Institute, and later encouraged me to take organic chemistry freshman

year of college. My mother has been similarly supportive, and helped me make my move to Ann

Arbor. I would also like to thank the many chemistry mentors that I have encountered over the

years, as they have guided me to where I am today. Early mentors include my high school

chemistry teacher Barbara Murphy, Roy Periana and the Periana lab, as well as Dan Weix, Michael

Prinsell, and the rest of the Weix lab. Pursuing a Ph.D. in chemistry can sometimes be a difficult

enterprise, and the positive experiences that they provided gave me the strength to pursue my

degree to completion.

I’ve had the pleasure of working with many collaborators over the years. Graduate students who

I’ve worked directly with include Alex Nett, Ellen Butler, Santiago Cañellas, Amie Frank on the

air-stable Ni(0) project, and Takunda Chazovachii, Danielle Fagnani, and Jessica Tami with

depolymerization. PIs I’ve worked with include Anne McNeil, Mark Banaszak Holl, and Neil

Marsh. The collaborations in the work reported here are enabled in part by the bounty of capable

students and PIs eager to collaborate that this department has to offer. Takunda Chazovachii in

particular has maintained a strong working relationship with me for a number of years.

The Procter and Gamble company has provided much of my financial support, and Dimitris Collias

Page 5: Computational Investigation of Chemical Reactions ...

iv

has worked with me on the P&G side of the poly(acrylic acid) project. P&G has also gifted me

with the opportunity to work there for a summer in 2017. Bruce Murch and Bill Laidig provided

me with a life-changing opportunity, and gave me the sense what industrial research is like.

The chemistry department has at the University of Michigan has been amazingly supportive over

the past 6 years, and I’m lucky to have the opportunity to work with two wonderful PIs. John

Montgomery was originally the sole chair of my committee, and he provided the support and

funding I needed for my first two years in the program. He was also more than accommodating

when it became clear that I was more suited for computational work. The investment in me from

him, Paul, and the department cannot be overstated. Paul Zimmerman, who became co-chair of

my committee in 2016, has been an excellent mentor. He has repeatedly amazed me with his deep

knowledge of disparate fields, and has been more than happy to provide valuable career advice.

Like John, Paul has shown a genuine care for his students, which has fostered in me a deep trust

in him. He was also instrumental in directing me towards my 2017 internship at Procter and

Gamble, as he recommended me for it based on confidence in me that I did not yet have in myself.

Finally, I would like to thank all of the friends that I’ve made in Ann Arbor over the years. My

coevals Aaron Proctor, Alonso Arguelles, and Grayson Ritch formed a core friend group that stood

the test of time. We put the effort in to regularly eat lunch and play games together, and it paid off

in the form of deep, lasting friendships. It seems unlikely to me that I’ll ever be able to forge

similar personal bonds like the ones formed among coevals in grad school. In addition to our group,

I’ve made many other friends, including Jacob Geri, Isaac Roles, Zach O’Keefe, and Cody Aldaz.

My experiences in grad school have been truly wonderful, and I’m forever grateful for the rich

social and intellectual life that Ann Arbor and the University of Michigan Department of

Chemistry have given me.

Page 6: Computational Investigation of Chemical Reactions ...

v

Table of Contents

Dedication ....................................................................................................................................... ii

Acknowledgements ........................................................................................................................ iii

List of Figures .............................................................................................................................. viii

List of Tables ................................................................................................................................ xv

Abstract ........................................................................................................................................ xvi

Chapter 1: Introduction ................................................................................................................... 1

1.1 Using Computational Chemistry to Evaluate Chemical Reactions ....................................... 1

1.2 The Potential Energy Surface ................................................................................................ 3

1.3 Using the Critical States of the PES to Derive Reaction Thermodynamics and Kinetics .... 4

1.4 The Computational Evaluation of Chemical States .............................................................. 8

1.5 Identifying Transition States Using the Growing-String Method ....................................... 10

1.6 Outlook ................................................................................................................................ 12

Chapter 2: Computational Investigation of a Nickel-Catalyzed Three-Component Coupling

Reaction ........................................................................................................................................ 13

2.1 Background ......................................................................................................................... 13

2.2 Mechanistic Ambiguity in the Reductive [3+2] Cycloaddition of Enoates and Alkynes ... 15

Page 7: Computational Investigation of Chemical Reactions ...

vi

2.3 Determining and Evaluating the Possible Mechanisms of the Three-Component Coupling of

an Alkyne, Aldehyde, and Enoate ............................................................................................. 17

2.4 Determining the Fates of Complexes VI-A, V-B, and IV-C ............................................... 22

2.5 Considering the Reactivity of α-Substituted Enoates .......................................................... 24

2.6 Summary and Conclusions .................................................................................................. 28

Chapter 3: Elucidation of the Activation Mechanism of Air-Stable Nickel(0) Catalysts............. 29

Some of the material presented in this chapter has been reported in ACS Catalysis.56 ............ 29

3.1 Motivation for the Development of Air-Stable Nickel Pre-Catalysts ................................. 29

3.2 Investigating a Diverse Set of Nickel Fumarate Complexes ............................................... 32

3.3 Overturning the Original Hypothesis that Catalyst Activation Occurs via Dissociation .... 34

3.4 Investigating the Activation Sequence of BAC and IMes Fumarate Complexes ............... 37

3.5 Preliminary Results Towards the Development of an Air-Stable Nickel(0) BAC Catalyst 42

3.6 Summary, Conclusions, and Outlook .................................................................................. 45

Chapter 4: Synergistic Effects Between Radical Attack and Applied Force in the

Depolymerization of Poly(Acrylic Acid)...................................................................................... 46

4.1 The Potential Energy Surface Under Applied Force ........................................................... 46

4.2 Plastic Recycling Using Ultrasonic Depolymerization ....................................................... 50

4.3 Evaluating the Impact of the Weak Bond Effect Using a Morse Potential ......................... 54

4.4 Determining the Effect of Radical Abstraction on the Transition State of Bond Scission . 58

4.5 Determining the Effective Tensile Strength of PAA With and Without Radical Attack .... 60

Page 8: Computational Investigation of Chemical Reactions ...

vii

4.6 Geometric Distortions Due to Applied Force on the Reactant and Transition State .......... 63

4.7 Rationalizing why PATH-1 is More Responsive to Force Than PATH-2 .......................... 68

4.8 Conclusions ......................................................................................................................... 71

Chapter 5: Conclusions and Final Thoughts ................................................................................. 73

5.1 Research Summary .............................................................................................................. 73

5.2 Possible Future Directions for the Material Studied in This Work ..................................... 79

5.3 On the Frailty of Human Understanding ............................................................................. 81

5.4 On the Symbiosis Between Theory and Experiment .......................................................... 82

Appendix ....................................................................................................................................... 84

Bibliography ................................................................................................................................. 99

Page 9: Computational Investigation of Chemical Reactions ...

viii

List of Figures

Figure 1-1. (Left) A 3-dimensional potential energy surface. (Right) The same PES, represented

using a single reaction coordinate. The energies of the individual states are listed on the right-hand

side. ................................................................................................................................................. 3

Figure 1-2. Equilibrium constants (1) between states A and B and (2) for a chemical reaction.

Where: Keq is the equilibrium constant, ΔG is the change in free energy, R is the gas constant, and

T is the absolute temperature of the reaction. ................................................................................. 5

Figure 1-3. The Eyring equation gives the rate of a chemical reaction, such as the transformation

of A to B. Where: krxn is the rate of the reaction, κ is the transmission coefficient, kb is Boltzmann’s

constant, T is the absolute temperature of the system, h is Planck’s constant, and R is the gas

constant. κ is traditionally considered to be set to 1. ...................................................................... 5

Figure 1-4. A potential energy surface for the transformation of A to E. The largest barrier process

involves B transforming into D via TS-C....................................................................................... 6

Figure 1-5. An example catalytic transformation of reactants to products. Two cycles of the

process are shown, with the intermediates of the second cycle denoted with an apostrophe. ........ 7

Figure 1-6. The process of finding a transition state using the single-ended growing string method.

....................................................................................................................................................... 11

Figure 2-1. (Top) The Ni-mediated [3+2] cycloaddition of alkynyl enal 2-1. (Bottom Left)

Proposed mechanism of cyclization. (Bottom Right) Reported crystal structure of metallacycle 2-

2, taken from ref. 49. ..................................................................................................................... 14

Page 10: Computational Investigation of Chemical Reactions ...

ix

Figure 2-2. The mechanism of [3+2] cyclization of enoates and alkynes, as proposed by

Montgomery (ref. 50). The formation of a linear side product when 2-3 is used suggests that

formation of a 7-membered metallacycle is part of the catalytic cycle. ....................................... 15

Figure 2-3 The mechanism of [3+2] cyclization of enoates and alkynes, as proposed by Ogoshi

(ref. 51). NMR characterization of a nickel-enolate complex suggests the intermediacy of a ketene-

containing species ......................................................................................................................... 17

Figure 2-4. (Top) Both Ogoshi and Montgomery propose intermediates that are vulnerable to

electrophilic attack. (Bottom) The optimized conditions of the three-component coupling between

an alkyne, aldehyde, and enoate. These conditions are used as the basis for the computational study

....................................................................................................................................................... 18

Figure 2-5. The four possible mechanisms investigated in the studied reaction. ........................ 19

Figure 2-6. Isomerization of metallacycle I to η3-bound III-A, and direct insertion through

pathway D. Pathway D is denoted in purple. Energies are given in kcal/mol, with enthalpies listed

in parentheses. Energies are given in kcal/mol, with enthalpies listed in parentheses. ................ 20

Figure 2-7. Comparison of the aldol-first (paths A, B) and ketene-first (path C) mechanisms. Path

A (black): inner-sphere aldol-first mechanism; Path B (blue): outer-sphere aldol-first mechanism;

Path C (red): ketene-first mechanism. Energies are given in kcal/mol, with enthalpies listed in

parentheses. ................................................................................................................................... 21

Figure 2-8. Aldol addition in pathway C. Energies are given in kcal/mol, with enthalpies listed in

parentheses. ................................................................................................................................... 22

Figure 2-9. Carbocyclization in paths A and B. Energies are given in kcal/mol, with enthalpies

listed in parentheses. ..................................................................................................................... 23

Page 11: Computational Investigation of Chemical Reactions ...

x

Figure 2-10. Formation of a linear side product in the studied reaction using an α-substituted

enoate. The existence of such a product implicates formation of a 7-membered metallacycle

(boxed). ......................................................................................................................................... 24

Figure 2-11. The phenoxide moiety moves close to the α position (highlighted in gray) after ketene

elimination. Increasing the steric hinderance at that position is expected to destabilize complex

III-C. ............................................................................................................................................. 25

Figure 2-12. The potential energy surface for the first few steps of the [3+2] cycloaddition between

an alpha substituted enoate and an alkyne. Energies are given in kcal/mol, with enthalpies listed in

parentheses. ................................................................................................................................... 26

Figure 2-13. Comparing the barrier to ketene elimination with isomerization to the 7-embered

metallacycle for both the α- and β-substituted enoates. ................................................................ 27

Figure 3-1. Use of a pre-catalyst dramatically improves the yield of 3-1. Crystal structure taken

from reference 78. ......................................................................................................................... 30

Figure 3-2. (Top) Air-stable Ni(II) catalysts require a transmetallation reagent. (Middle) Tradeoffs

in the development of Ni(0) pre-catalysts. (Bottom) Molecular orbital diagram for the π

backbonding interaction. ............................................................................................................... 31

Figure 3-3. (Top) Different fumarate complexes were synthesized. (Bottom) The different

complexes were tested for their ability to catalyze a reductive coupling reaction. ...................... 33

Figure 3-4. Other complexes of fumarate A are not competent in reductive couplings. ............. 34

Figure 3-5. The mechanism of the reductive coupling reaction. Originally ligand dissociation was

believed to be the means of catalyst activation. ............................................................................ 34

Page 12: Computational Investigation of Chemical Reactions ...

xi

Figure 3-6. The relationship between the free energy of ligand exchange and the yield of 3-9 in

reductive coupling attempts with the corresponding catalyst. Computational details are included

in the appendix. ............................................................................................................................. 35

Figure 3-7. The isolation of products 3-12 and 3-13 from catalyst 3-8-A. .................................. 36

Figure 3-8. Possible activation mechanisms for complexes of fumarate A. ................................ 37

Figure 3-9. The initial steps of activation for IMes complex 3-8-A (blue and turquoise) and BAC

complex 3-10-A (red and pink). Dark colors (red, blue and black) represent aldol-first path A.

Light colors (pink, turquoise, and gray) represent ketene-first path B. Energies are given in

kcal/mol, with enthalpies listed in parentheses. Computational details are included in the appendix.

....................................................................................................................................................... 38

Figure 3-10. Pathway A of the activation of IMes catalyst 3-8-A, part 1. Energies are given in

kcal/mol, with enthalpies listed in parentheses. Computational details are included in the appendix.

....................................................................................................................................................... 40

Figure 3-11. Pathway A of the activation of IMes catalyst 3-8-A, part 2. Path Z details the possible

ester reduction reaction, and is shown in seafoam green. Energies are given in kcal/mol, with

enthalpies listed in parentheses. Computational details are included in the appendix. ................ 41

Figure 3-12. (Left) Complex BAC-IV-B can rearrange into carbonyl bound complex BAC-V-B.

(Right) A three-dimensional representation of complex BAC-V-B. ............................................ 42

Figure 3-13. Preliminary efforts towards developing a BAC fumarate catalyst. Fumarates that

favor metallacycle isomerization over ketene elimination are shown in blue. Fumarates where

metallacycle isomerization is within 1 kcal/mol or less than ketene elimination are shown in purple.

Listed energies are free energies in kcal/mol, and are relative to BAC-I (Figure 3-9).

Computational details are included in the appendix. .................................................................... 43

Page 13: Computational Investigation of Chemical Reactions ...

xii

Figure 4-1. The additive effect of applied force on the Morse potential. The distortion due to

applied force is greater at larger C-C bond distances. .................................................................. 46

Figure 4-2. Estimating the tensile of a C-C bond using the Morse method. ................................ 47

Figure 4-3. The expected effects of force according to the tilted potential energy surface model.

....................................................................................................................................................... 48

Figure 4-4. Recycling of polymers is enabled by ultrasonic depolymerization. .......................... 51

Figure 4-5. This chapter explores the synergistic interplay between radical attack and tensile force.

....................................................................................................................................................... 53

Figure 4-6. Model systems used in this study: Degradation of PAA through force alone (PATH-

1), and force in conjunction with radical attack (PATH-2). ......................................................... 55

Figure 4-7. Flowchart for evaluating tensile strength using the Morse method. The complete

procedure is described in the appendix. ........................................................................................ 55

Figure 4-8. Example three-dimensional representations of the tetramer (Right) and dimer+dimer

(Left) systems used for this study. These represent two of many snapshots used for this study. . 56

Figure 4-9. The effect of hydrogen atom abstraction on the tensile strength poly(acrylic acid).

Computational details are included in the appendix. .................................................................... 57

Figure 4-10. The relationship between enthalpy of activation and applied force on the polymer

chain. Computational details are included in the appendix. ......................................................... 59

Figure 4-11. Derivation of a relationship between enthalpy of activation (ΔH‡) and reaction half-

life (t1/2). Where: kb is Boltzmann’s constant, T is the reaction temperature, h is Planck’s constant,

R is the gas constant, ΔH‡ and ΔS‡ are the enthalpy and entropy of activation, ΔHpolym and ΔSpolym

are the enthalpy and entropy of polymerization, and Tc is the ceiling temperature of

polymerization. ............................................................................................................................. 61

Page 14: Computational Investigation of Chemical Reactions ...

xiii

Figure 4-12. A contour plot of the effective tensile strength (E)FT of 1 and 2 assuming a Tc value

of 200 °C. ...................................................................................................................................... 62

Figure 4-13. Comparison of key geometries, which suggest that PATH-1 initial and TS structures

converge towards one another as forces increase, but PATH-2 structures do not. Explicit water

molecules are removed for clarity................................................................................................. 63

Figure 4-14. The effects of applied force on the end-to-end lengths in PATH-1 and PATH-2. 64

Figure 4-15. The effects of applied force on the scissile bond distances in PATH-1 and PATH-2.

....................................................................................................................................................... 66

Figure 4-16. The energetic contribution from applied force (ΔEAF) observed in PATH-1 and

PATH-2. The decrease in ΔH‡ with force in PATH-2 can be explained though the steady increase

in ΔEAF. ......................................................................................................................................... 67

Figure 4-17. Comparing the (relative) end-to-end and scissile bond distances of the starting and

TS structures in PATH-1 and PATH-2. The average end-to-end and scissile bond distances were

calculated within each reaction path. ............................................................................................ 68

Figure 4-18. Differences in the PES curvature of PATH-1 and PATH-2. ................................. 69

Figure 4-19. (Top) How the Morse and quartic potentials behave under applied force (Bottom

Left) Comparing the predicted TS energies of PATH-1 and PATH-2 using the example Morse

and quartic functions. The ratio of the change in energy with force (ESlope1 vs ESlope2) is very

close to the atomistic model. (Bottom Right) Comparing the displacement of the location of the

TS using the example Morse and quartic functions. The ratio of the change in displacement with

force at the last five data points (DispSlope1 vs DispSlope2) is consistent with the atomistic model.

....................................................................................................................................................... 70

Page 15: Computational Investigation of Chemical Reactions ...

xiv

Figure 5-1. A summary of Chapter 2. Two mechanisms for the three-component coupling reaction

were found to be plausible. The selectivity for a specific mechanism appears to be affected by the

placement of a methyl group at different positions of the enoate. ................................................ 75

Figure 5-2. (A) Original design strategy to develop an air-stable Ni(0) pre-catalyst. (B) No

Correlation between reaction yield and the free energy of ligand exchange was observed. (C) The

absence of such a correlation suggests that a fumarate consumption reaction must be active. .... 76

Figure 5-3. The insights gained from Chapter 2 explain why complex 3-8-A is active, but complex

3-10-A is not. The two complexes undergo activation through different mechanisms. 3-8-A goes

through an aldol-first mechanism, allowing for catalyst activation, but 3-10-A undergoes a ketene-

first activation, resulting in a stable ester-ligated complex. .......................................................... 77

Figure 5-4. A summary of Chapter 4. (A) Radical abstraction imparts a significant change in the

enthalpy of bond scission. (B) The barrier for bond scission in PATH-1 and PATH-2. (C)

Differences in the geometries of key states in PATH-1 and PATH-2. (D) We can explain the

differences in behavior between PATH-1 and PATH-2 to be due to the changes in the curvature

of the potential energy surfaces of the two processes. .................................................................. 78

Figure A-1. Model PAA systems used in this study. ................................................................... 88

Figure A-2. Flowchart for QM/MM optimization and bond scission. ......................................... 89

Figure A-3. Points where force is applied on compounds 1 and 2. The scissile bonds are shown in

red. ................................................................................................................................................ 91

Figure A-4. Flowchart for identifying the transition state of bond scission for compound 1. ..... 92

Figure A-5. Flowchart for identifying the transition state of bond scission for compound 2. ..... 92

Page 16: Computational Investigation of Chemical Reactions ...

xv

List of Tables

Table 4-1.The effective tensile strength of 1 and 2 calculated under different assumptions. ...... 61

Table A-1. Energetic values used to calculate relative energies in the activation pathway of IMes

catalyst 3-8-A. ............................................................................................................................... 85

Table A-2. Energetic values used to calculate relative energies in the activation pathway of BAC

catalyst 3-10-A. ............................................................................................................................. 86

Table A-3. Energetic values of small molecules used in the calculations of the activation pathways

of both 3-8-A and 3-10-A. ............................................................................................................ 86

Table A-4. Energetic values used in the comparison of ketene elimination with metallacycle

isomerization of BAC complexes using different fumarates. ....................................................... 87

Table A-5. Average calculated bond dissociation energy values in PATH-1 and PATH-2. ...... 95

Table A-6. Average calculated force constants in PATH-1 and PATH-2. .................................. 95

Table A-7. The average barriers for bond scission in PATH-1 and PATH-2. ............................ 95

Table A-8. The average end-to-end and scissile bond distances in 1, 2, TS-1, and TS-2. .......... 96

Table A-9. Parameters used for mathematical modeling in Chapter 4. ....................................... 97

Table A-10.The change in the transition state energy of the mathematical models of PATH-1 and

PATH-2. ....................................................................................................................................... 97

Table A-11. Displacement of the position of the transition state in the mathematical models of

PATH-1 and PATH-2. ................................................................................................................. 98

Page 17: Computational Investigation of Chemical Reactions ...

xvi

Abstract

Computational chemistry has emerged as a powerful tool in identifying the key chemical states

that define the kinetics, thermodynamics, and the various selectivities of a chemical reaction. It

allows for a greater understanding of chemical processes by providing insights that would be

difficult to achieve through direct experimentation alone. Herein are described three examples

where computational analysis is used to provide insight into chemical reactions.

In Chapter 1, some fundamentals of computational research in chemistry is described. Topics

include the potential energy surface, the computational methods used for energy calculation, and

transition state finding using the growing string method. The tools described in this chapter are

used in the subsequent chapters.

Chapter 2 investigates the mechanism of a nickel-catalyzed [3+2] alkylative cycloaddition. Prior

experimental work has suggested multiple possible mechanisms, which are interrogated here

computationally. Two distinct mechanisms are found to be in a precarious balance with each other,

and small perturbations in the substrates used in the reaction are found to tilt the balance towards

one mechanism over the other. The findings from this chapter have implications for the mechanism

of [3+2] alkylative cycloadditions as a whole, including the activation process detailed in Chapter

3.

In Chapter 3, the activation process of air-stable nickel fumarate catalysts is studied. The original

hypothesis of fumarate dissociation as a mechanism of activation was overturned after

Page 18: Computational Investigation of Chemical Reactions ...

xvii

computational analysis determined a fumarate consumption event must occur instead. These

findings inspired subsequent experimental work that isolated the products of fumarate dissociation,

and found that catalyst activation goes through a similar [3+2] alkylative cycloaddition detailed in

Chapter 2. Through this lens, the mechanisms of activation of active IMes catalyst and inactive

BAC catalyst are evaluated computationally, to provide a rationale for why one catalyst works and

the other does not. Using these insights, preliminary investigations towards improving the BAC

catalyst are performed.

Chapter 4 details the effect of radical attack on the tensile strength of poly(acrylic acid) (PAA).

Two different methods are used to determine how radical attack weakens the tensile strength of

the polymer backbone. Additionally, the effect of force in the radical-free and radical-abstracted

cases on the geometry of the starting structures and transition state of the species involved in bond

scission is observed. The different behaviors between the two regimes are attributed to the

differences in the curvature of their potential energy surfaces. The findings of this chapter carry

implications for the development of new depolymerization reactions.

Finally, Chapter 5 summarizes the work, and provides final thoughts on future directions. Elements

of this work are compared to the Gettier problem, and the symbiotic relationship between theorists

and experimentalists is discussed.

Page 19: Computational Investigation of Chemical Reactions ...

1

Chapter 1: Introduction

1.1 Using Computational Chemistry to Evaluate Chemical Reactions

In considering the development of a new chemical reaction, two conditions must hold true. Firstly,

the reaction must be thermodynamically feasible. This condition is satisfied if the entire chemical

process is exergonic, which means that the free energy of the final product is lower than the free

energy of the starting material. Secondly, the reaction must be kinetically feasible. While any

exergonic reaction is guaranteed to happen eventually, a kinetically feasible reaction will occur at

a rate that is useful to humans, and will also outcompete other potential pathways. The rate at

which a chemical reaction occurs is controlled by the energy of its highest energy state that occurs

during the chemical transformation.

To evaluate the kinetics and thermodynamics of a chemical reaction, therefore, all that is needed

to be done is to evaluate the energy of a few certain states. The challenge in computational

chemistry, then, is identifying the states that are relevant over the course of the reaction, and

accurately evaluating them. Fortunately, quantum chemical techniques such as density functional

theory (DFT) have emerged as a powerful tool to evaluate the energies of chemical structures.1

Additionally, recent development of transition-state finding and conformer-generating tools has

given computational chemists the ability to identify relevant chemical states at an unprecedented

pace.2

Page 20: Computational Investigation of Chemical Reactions ...

2

By leveraging these advancements, computational chemists are now in a position to provide useful

predictions, at a pace fast enough to keep up with experimental advancements. Chemical reactions

can often be challenging to characterize experimentally, and computational chemistry has the

advantage of being able to directly study the putative intermediates of a reaction. This contrasts

with the majority of experimental research in reaction development, which often focuses on the

inputs and outputs of a reaction, and can sometimes struggle to identify key intermediates.

This work describes a number of cases where computational chemistry is used to explain and

predict experimental results. Chapter 2 details the mechanism of a nickel-catalyzed three-

component coupling reaction. Previous work on related [3+2] reductive cycloadditions gave

conflicting evidence on what the active mechanism of the transformation could be, and the addition

of computational analysis provides clarity to the murky experimental picture. In Chapter 3, the

activation sequence of air-stable nickel(0) catalysts are examined. This chapter showcases the

ability of computational chemistry to provide useful predictions for the experimentalist, and uses

the knowledge gained from Chapter 2 as a framework for understanding catalyst activation. And

finally, in Chapter 4, the effects of hydrogen atom abstraction in the mechanochemical degradation

of poly(acrylic acid) are explored. Here, recently developed tools for transition state finding under

force are used, which allow for analysis of the differences in structure between bond scission under

force in the presence and absence of a radical on the polymer backbone.

The remainder of this chapter will provide a useful background for understanding the

computational analysis of chemical reactions. Topics to discuss include: defining the potential

energy surface, identifying and utilizing the key chemical states necessary to describe the

thermodynamics and kinetics of a reaction, computational methods to evaluate chemical states,

and modern transition state finding methods.

Page 21: Computational Investigation of Chemical Reactions ...

3

1.2 The Potential Energy Surface

Figure 1-1. (Left) A 3-dimensional potential energy surface. (Right) The same PES, represented using a single

reaction coordinate. The energies of the individual states are listed on the right-hand side.

For a given chemical system, there exists a multitude of possible arrangements of the atoms

involved in the system. If we imagine all of the arrangements, we can construct a hypersurface

existing in approximately 3N dimensions, with N being the number of atoms in the system. This

hypersurface is generally referred to as the “potential energy surface” (PES), and its contours are

what determines the mechanisms of chemical transformations. A simplified multi-dimensional

representation of such a surface is given on the left-hand side of Figure 1-1. In Figure 1-1, points

1, 2, and 3 are local minima on the surface, and which means they also represent chemical

intermediates. Intermediate 1 is capable of transitioning to 2 through the path containing TS-1,

and intermediate 2 can then transform into 3 through the path containing TS-2. Points TS-1 and

TS-2 are notable, as they are at saddle-points on the PES. The technical definition of a saddle point

is a point where the gradient equals zero, but is neither a local maximum or minimum. In the

context of analyzing the PES of a chemical system, saddle points are the highest energy point in

the lowest energy path between two local minima. The most probable path between two

Page 22: Computational Investigation of Chemical Reactions ...

4

intermediates will go through the saddle point, which in chemical terms is referred to as a transition

state, abbreviated as “TS” in this text.

While the PES provides a complete description of all possible atomic arrangements for a given

system, in practice, it is more useful to pictorially describe the PES along a single axis, as

represented on the right-hand side of Figure 1-1. This representation makes it easier to see the

energies of the different states involves, and simplifies the interpretation of the PES. By focusing

only on the critical states (intermediates and transition states), we can more easily determine the

kinetics and thermodynamics of a given chemical reaction.

1.3 Using the Critical States of the PES to Derive Reaction Thermodynamics and Kinetics

Determining the thermodynamics of a chemical reaction is a straightforward exercise. In looking

at the equilibrium between two states, the equilibrium constant is given by equation 1 (Figure 1-2).

Notably, as the difference in free energy (∆𝐺) grows larger, the equilibrium constant (𝐾𝑒𝑞) changes

at an exponential rate. In looking a unimolecular transformation, where species A transforms into

species B, the equilibrium constant is the expected ratio between the two states. Using the

relationship in equation 1, we can evaluate the thermodynamics of an entire chemical reaction

using equation 2 (Figure 1-2), where multiple reactants (R1, R2,…) are converted into multiple

products (P1, P2, …). Based off of equation 2, we can see that when a reaction is exergonic (∆𝐺 <

0), 𝐾𝑒𝑞 becomes larger than 1, indicating that if the reaction mixture were allowed to equilibrate,

the majority of the material would transform into the set of products. Additionally, given the

exponential nature of the equilibrium constant, a ∆𝐺 value of merely -10 kcal/mol would be

expected to equilibrate until the ratio of products to reactants exceeds 1,000,000 to 1.

Page 23: Computational Investigation of Chemical Reactions ...

5

Figure 1-2. Equilibrium constants (1) between states A and B and (2) for a chemical reaction. Where: Keq is the

equilibrium constant, ΔG is the change in free energy, R is the gas constant, and T is the absolute temperature of the

reaction.

While evaluating the thermodynamics a reaction requires the analysis of only the start and end

states, the kinetics of reaction is more difficult to compute. The rate at which a chemical reaction

progresses is given by the Eyring equation, shown in Figure 1-3. Under transition-state theory, the

transition state is considered to be in quasi-equilibrium with the reactants. Thus, as the Eyring

equation is based off of the equations for thermodynamic equilibrium, it shares many similar

features. In particular, under the Eyring equation, the difference in free energy between the reactant

and the transition state (∆𝐺‡) affects the rate at which the reaction can happen. As ∆𝐺‡ increases,

the rate of transformation decreases exponentially.

Figure 1-3. The Eyring equation gives the rate of a chemical reaction, such as the transformation of A to B. Where:

krxn is the rate of the reaction, κ is the transmission coefficient, kb is Boltzmann’s constant, T is the absolute

temperature of the system, h is Planck’s constant, and R is the gas constant. κ is traditionally considered to be set to

1.

However, a typical chemical reaction tends to involve multiple intermediates, with multiple

transition states where one intermediate transforms into another. In the multistep case, the time

Page 24: Computational Investigation of Chemical Reactions ...

6

that it takes for a unit of reactant to transform into a unit of product would be expected to be a

combination of the total time that it takes for all steps in the reaction to occur.3 However, this

picture can be greatly simplified by taking advantage of the exponential nature of the Eyring

equation. At room temperature, for every additional 1 kcal/mol increase in ∆𝐺‡, the corresponding

reaction takes about 5.4 times longer to occur. Thus, the largest overall ∆𝐺‡ ends up accounting

for the vast majority of the time needed for a chemical transformation to happen.3

Figure 1-4. A potential energy surface for the transformation of A to E. The largest barrier process involves B

transforming into D via TS-C.

Searching for the largest overall barrier is not a simple case of evaluating the energy of a transition

state relative to its prior intermediate. For instance, in Figure 1-4, in the transformation of A to E,

the largest barrier for the reaction is 16.0 units, measured between intermediate B and transition

state TS-C. The overall barrier is calculated to be 16.0 units as intermediate C is higher in energy

than B. That means, thermodynamically, only a small fraction of B will convert to C, and then, of

that small fraction, some of that material will convert to D through transition state TS-C. By

measuring the barrier from B to TS-C, we take into account the fact that conversion to C is a

thermodynamically unfavorable process. And as the Eyring equation is derived from the equations

Page 25: Computational Investigation of Chemical Reactions ...

7

in Figure 1-2, we can simply add the difference in energy between B and C to the barrier from C

to TS-C to derive an overall barrier for the transformation.3

A similar analysis can be performed on catalytic reactions.4–6 But, since catalytic reactions are by

nature cyclical, it’s possible that the catalytic cycle can initially form a very stable complex, and

then subsequently fail to “turn over”. The largest barrier can be found by looking at two contiguous

catalytic cycles, with the second cycle offset by the difference in energy between the reactants and

the products. With the new potential energy surface in hand, searching for the highest barrier can

proceed as normal. An example of this process is shown in Figure 1-5. In this example, the highest

barrier process involves intermediate B turning over the catalyst via TS-A’. Without considering

a second catalytic cycle, it would appear that the highest barrier process would be intermediate A

transforming into intermediate B, leading to an erroneous lower barrier, and a mistaken belief

about which states in the catalytic cycle have the greatest effect on catalyst turnover.

Figure 1-5. An example catalytic transformation of reactants to products. Two cycles of the process are shown, with

the intermediates of the second cycle denoted with an apostrophe.

Page 26: Computational Investigation of Chemical Reactions ...

8

1.4 The Computational Evaluation of Chemical States

Energy evaluation is a necessary component in the construction of a potential energy surface. The

main method used in this work is density functional theory (DFT). DFT is considered to have been

developed in 1964,1 when Hohenberg and Kohn proved that kinetic and potential energies of any

N-electron system can be described by a function of electron density (a so-called “density

functional”), implying that one could perfectly predict the properties of a system, if a perfect

density functional was in hand.7 The following year, Kohn and Sham demonstrated that a density

functional that treats all of the electrons as non-interacting accounts for the bulk of the total energy,

meaning that only the small effect of electron interaction, termed the “exchange-correlation

energy” needs to be determined to obtain an exact result.8

By avoiding the direct computation of electron-electron interactions, DFT scales at a relatively

low cost (N3 , where N is the number of electrons), which compares favorably with wavefunction

based methods (N4-10).9 While in principle, DFT can deliver perfect accuracy with a perfect

functional that describes the exchange-correlation energy, in practice, no such functional exists,

and the exchange-correlation energy is approximated. The approaches towards approximating the

exchange-correlation energy are diverse and numerous, and this had led to the development of a

large menagerie DFT functionals, with differing accuracies and computational costs.10 In this work

the ωB97X-D311 functional is used in Chapters 2 and 3, and the B3LYP12 functional is used in

Chapter 4. While the errors calculated DFT functionals are dependent on the benchmarking set

used, they provide a useful sense of the accuracy of the functional. In interpreting the

computational results of this work, an error of approximately 2 kcal/mol for ωB97X-D3, and

approximately 5 kcal/mol for B3LYP, should be considered.13

Page 27: Computational Investigation of Chemical Reactions ...

9

Another consideration in the accuracy of DFT is the choice of basis set for the calculation. While

a DFT functional ultimately defines the density of electrons a given chemical system, the basis set

describes the orbitals within which the electrons reside. The choice of which orbitals to define, and

how to define them, is another factor that can influence the cost and accuracy of a calculation. In

general, “double zeta” basis sets (where two functions are used to describe each atomic orbital)

are used for processes such as geometry optimization or transition-state finding, where many

iterative energy and gradient calculations are performed, and the more computationally expensive

“triple zeta” basis sets (three functions per atomic orbital) are used for accurate energy calculations

on an optimized structure.14

Ab initio methods such as DFT only require the choice of basis set, functional, and the cartesian

coordinates for the atoms in the chemical system to be studied. Their simplicity makes them well

suited to investigate structures where features such as bond lengths, angles, and torsions are

ambiguous, such as in most transition states or metal complexes. However, arriving at those

features from first principles is computationally taxing, and in our hands, DFT calculations on

systems larger than 200 atoms are too slow to be of use. In molecular mechanics (MM), the

parameters that can define features such as bond lengths, angles, and torsions are determined

beforehand, often fit to quantum chemical calculations. Using these parameters to calculate the

energy of the system results in a speed up of multiple orders of magnitude. In general, molecular

mechanics is attractive for lager systems, and is often used to simulate proteins or molecules in the

condensed phase.15

As molecular mechanics is a good choice for the simulation of large condensed phases, and

quantum methods such as DFT is a good choice for the simulation of transition states, a method

that uses both quantum and molecular mechanical techniques may be best suited for the study of

Page 28: Computational Investigation of Chemical Reactions ...

10

transition states within a large system. In QM/MM, a portion of a chemical system is treated using

a quantum method such as DFT, MP2, or semi-empirical methods, and the remainder of the system

is treated using molecular mechanics.16 In Chapter 4 of this work, a QM/MM system is used with

CHARMM2717,18 molecular mechanics parameters, in conjunction with DFT using the B3LYP

functional with the 6-31G* basis, using an electronic embedding system.19

1.5 Identifying Transition States Using the Growing-String Method

While the structure of chemical intermediates only requires following the analytic first derivative

to find a local minimum,20 transition state finding is more difficult, as it is neither a local maximum

or minimum on the PES, so a strategy of only following the gradient cannot be used. Fortunately,

a number of tools have been developed for transition state searching. These include the linear

synchronous transit method,21 the nudged elastic band method,22,23 and many others. In the

Zimmerman group, a growing string method (GSM) using an internal coordinate system has been

developed,24–26 and is the method of choice in this work for the purposes of identifying transition

states. The chief advantages of using GSM is that it allows for the user to be ignorant of the

potential structure of the transition state, and, in the case of single-ended GSM, also be ignorant

of the potential structure of the product.

The process for single-ended GSM26 is detailed in Figure 1-6. In the example reaction, the SN2

addition of thiolate to methyl bromide, the user defines the bonds that they would like to add and

break (Top Left). The add and break moves are then interpreted to define the R-P (reactant-

product) tangent, a vector on the potential energy surface that dictates the general direction to find

a transition state (Top Right). In the direction of the R-P tangent, a new structure (referred to as a

node) can be generated (Figure 1-6, 1). This node can then be optimized in all directions except

along the R-P tangent (2). From the optimized node, a second node can be generated along the

Page 29: Computational Investigation of Chemical Reactions ...

11

Figure 1-6. The process of finding a transition state using the single-ended growing string method.

Page 30: Computational Investigation of Chemical Reactions ...

12

R-P tangent (3), and then optimized again (4), creating a string of nodes. This process can be

repeated, until the nodes begin to decrease in energy, indicating that a saddle point has been crossed

(5). Finally, to identify the transition state, the nodes along the growing string are optimized again,

and then the highest-energy node is optimized to a transition state using a climbing image search

(6).25

1.6 Outlook

Despite rough complexity and high dimensionality of the PES, a reaction can be evaluated using

only a handful of points on the hypersurface. The thermodynamics of the entire process can be

determined by comparing the free energy of all of the reactants compared to all of the products.

The kinetics of a reaction can be determined by the observing the largest possible climb in energy

between an intermediate and a transition state that occurs in the reaction. Even though in theory,

the energetics of only four states need to be evaluated to have an understanding of the kinetics and

thermodynamics of a reaction, in practice, more states are typically evaluated, if only to be certain

that those processes are possible. These techniques and concepts will be used in the coming

chapters to evaluate the chemical process of a three-component coupling reaction, the activation

sequence of an air-stable nickel(0) catalyst, and the depolymerization of poly(acrylic acid)

Page 31: Computational Investigation of Chemical Reactions ...

13

Chapter 2: Computational Investigation of a Nickel-Catalyzed

Three-Component Coupling Reaction

The content in this chapter has been published in the Journal of Organic Chemistry.27

2.1 Background

Five-membered carbocycles are a common structural motif in biology, which has resulted in the

development of a rich body of literature on their synthesis.28–33 Within that body of research,

countless strategies for 5-membered ring formation have been developed, including the Nazarov

cyclization,32 ring-expansion reactions,34 ring-closing metathesis,35 radical cyclizations,36

intramolecular nucleophilic attack,33 cross-couplings,37 or cycloisomerization reactions.33 In

addition to these listed methods, cycloaddition reactions, in which a ring is formed from multiple

separate components, have also been developed. These can include [3+2] cycloadditions, [2+2+1]

cycloadditions, and [4+1] cycloadditions.33

[3+2] cycloadditions in particular have developed into a powerful synthetic method, due to their

potential for chemo-, regio-, diasterio- and enantioselectivity.28 These transformations can be

accomplished through the use of donor-acceptor cyclopropanes,38 vinyl carbenoid additions,28

allyl- or allenylsilane additions,39 or ylide additions.40 Reductive metal-promoted [3+2]

cycloadditions have also represented a viable strategy towards cyclopentane formation. A number

of processes have been developed to couple readily available reagents such as alkynes and α,β-

unsaturated carbonyl compounds. In such processes, stoichiometric amounts of nickel,41

Page 32: Computational Investigation of Chemical Reactions ...

14

titanium,42,43 iron,44,45 or cobalt46 metallacycles are formed, which can then then collapse into a 5-

membered ring after protonation or alkylation (Figure 2-1).

Figure 2-1. (Top) The Ni-mediated [3+2] cycloaddition of alkynyl enal 2-1. (Bottom Left) Proposed mechanism of

cyclization. (Bottom Right) Reported crystal structure of metallacycle 2-2, taken from ref. 49.

The Montgomery group has a long history of using nickel metallacycles in cycloaddition

reactions.41,47 The group first reported a nickel-mediated [3+2] reductive cyclization in 2000,48

combining alkynyl enal 2-1 with stochiometric Ni(COD)2 and TMEDA, to create 7-membered

metallacycle 2-2, which can then be quenched with water, methyl iodide, or benzaldehyde to create

Page 33: Computational Investigation of Chemical Reactions ...

15

a cyclopentenone (Figure 2-1). This transformation was though to occur via initial nucleophilic

attack by the enolate moiety in 2-2, followed by subsequent nucleophilic attack on the resulting

aldehyde by the nickel-vinyl species. Metallacycle 2-2 was later directly isolated, and its structure

confirmed by X-ray crystallography.49

2.2 Mechanistic Ambiguity in the Reductive [3+2] Cycloaddition of Enoates and Alkynes

Figure 2-2. The mechanism of [3+2] cyclization of enoates and alkynes, as proposed by Montgomery (ref. 50). The

formation of a linear side product when 2-3 is used suggests that formation of a 7-membered metallacycle is part of

the catalytic cycle.

Page 34: Computational Investigation of Chemical Reactions ...

16

The stoichiometric reactions detailed in Figure 2-1 paved the way for methods catalytic in nickel

to be developed. In 2011, Montgomery50 reported catalytic, intramolecular [3+2] cycloaddition

reactions of enals with alkynes and of enoates with alkynes. In a separate, concurrent report,

Ogoshi51 also detailed the cycloaddition of enoates and alkynes, using isopropanol as a reductant.

Remarkably, even though both reports detail similar transformation, and both use a nickel catalyst

with a strong σ-donor ligand, Montgomery and Ogoshi attribute the formation of the [3+2]

cycloadduct to separate mechanisms, and both authors provide compelling evidence to support

their mechanistic proposals.

In Montgomery’s publication, the proposed mechanism of cyclopentenone formation involves the

cyclization of the enoate and alkyne, and then subsequent isomerization to a 7-membered

metallacycle (Figure 2-2).50 The resulting metallacycle can be protonated to form a nickel species

with a π-bound carbonyl, and then undergo carbocyclization, followed by subsequent alkoxide

extrusion to yield the desired cyclopentenone product. Support for this mechanism stems from the

observation of linear side products from the attempted cycloadditions of enoate 2-3, where the

ester moiety is kept intact. These observed side products are comparable to previous work52,53 from

the group in which enals and enones are reductively coupled with alkynes to create similar

products. Their formation is attributed to the intermediacy of a 7-membered metallacycle, similar

to the previously reported 2-2.

However, in Ogoshi’s report, a different mechanistic picture is painted (Figure 2-3).51 In that

report, it was posited that after the enoate and alkyne cyclize, rather than isomerize to a 7-

membered metallacycle, the complex instead undergoes phenoxide elimination to form a ketene

complex. The ketene complex can then undergo carbocyclization to yield a nickel enolate species

that can then be protonated off. Catalyst regeneration occurs via β-hydride elimination of the

Page 35: Computational Investigation of Chemical Reactions ...

17

resulting nickel-isopropoxide complex. To support this mechanistic proposal, Ogoshi also reports

the NMR characterization of the nickel enolate species with an IPr ligand, which is formed quickly

from starting materials in the absence of isopropanol, and quickly decays into product once

isopropanol is added.

Figure 2-3 The mechanism of [3+2] cyclization of enoates and alkynes, as proposed by Ogoshi (ref. 51). NMR

characterization of a nickel-enolate complex suggests the intermediacy of a ketene-containing species

2.3 Determining and Evaluating the Possible Mechanisms of the Three-Component Coupling

of an Alkyne, Aldehyde, and Enoate

As both Montgomery and Ogoshi provide concrete support for their proposed mechanisms, it

seems unlikely that only one mechanism is active in the [3+2] cycloaddition between an enoate

Page 36: Computational Investigation of Chemical Reactions ...

18

alkyne. Rather, it is more likely that multiple mechanisms are possible, and small changes in the

reactant conditions can push the reaction to favor one mechanism over another. However, despite

the difference between the two proposed mechanisms, in both cases, catalyst turnover is enabled

by the protonation of a nickel enolate species with an alcohol. If the alcohol was removed from

the reaction conditions, and replaced by an electrophilic carbon, it could be possible to form an

additional carbon-carbon bond in the process (Figure 2-4, top).

Figure 2-4. (Top) Both Ogoshi and Montgomery propose intermediates that are vulnerable to electrophilic attack.

(Bottom) The optimized conditions of the three-component coupling between an alkyne, aldehyde, and enoate. These

conditions are used as the basis for the computational study

Through careful optimization of reaction conditions, Montgomery group student Aireal Jenkins

was able to realize a catalytic, three-component coupling reaction by incorporating an aldol

reaction in with the [3+2] reductive cycloaddition of an enoate and alkyne (Figure 2-4, bottom).

With a working reaction in hand, it was at this point that computational analysis was requested.

Page 37: Computational Investigation of Chemical Reactions ...

19

The goals of the computational study were to clarify the mechanism through which the three-

component coupling occurs. While multiple sets of conditions were developed for the reaction,

including using either phosphine ligand PBu3 or NHC (N-heterocyclic carbene) ligand IMes, the

conditions listed in Figure 2-4 were chosen for computational study due to the popularity of NHC

ligands in the more recent reductive coupling work by the Montgomery group.54

Figure 2-5. The four possible mechanisms investigated in the studied reaction.

Based on the mechanisms proposed by Montgomery50 and Ogoshi51 in their respective reports, as

well as from reviewer input in the publication of the data in this chapter, four possible mechanisms

for the three-component coupling can be envisioned (Figure 2-5). Pathways A and B are based on

the mechanistic proposal of Montgomery, with the intermediacy of a 7-membered metallacycle.

In these so-called “aldol-first” pathways, the aldol addition occurs prior to carbocyclization. In

pathway A, the unit of aldehyde coordinates directly to the metal center, allowing for an inner-

Page 38: Computational Investigation of Chemical Reactions ...

20

sphere aldol addition. In pathway B, the aldol addition occurs in an outer-sphere process without

any prior aldehyde coordination. Pathway C is a “ketene first” mechanism based off of the

mechanistic proposal of Ogoshi, where carbocyclization occurs prior to aldol addition. And in the

lass possible mechanism, pathway D, a coordinated aldehyde inserts directly into the 5-membered

metallacycle, based off of reviewer comments during publication review.

Figure 2-6. Isomerization of metallacycle I to η3-bound III-A, and direct insertion through pathway D. Pathway D is

denoted in purple. Energies are given in kcal/mol, with enthalpies listed in parentheses. Energies are given in

kcal/mol, with enthalpies listed in parentheses.

As all four investigated mechanisms begin with the oxidative cyclization of the enoate and alkyne,

metallacycle I (Figure 2-6) is used as the reference structure. In order to isomerize to the 7-

membered metallacycle, I must first isomerize to η3 intermediate III-A, via rotation (TS-I) to

isomer II, followed by carbonyl binding (TS-II-A) to yield III-A. Additionally, it is also possible

Page 39: Computational Investigation of Chemical Reactions ...

21

for benzaldehyde to coordinate to metallacycle I to yield complex II-D, which can then directly

insert (TS-II-D), to create tetracoordinate species VI-A.

Figure 2-7. Comparison of the aldol-first (paths A, B) and ketene-first (path C) mechanisms. Path A (black): inner-

sphere aldol-first mechanism; Path B (blue): outer-sphere aldol-first mechanism; Path C (red): ketene-first

mechanism. Energies are given in kcal/mol, with enthalpies listed in parentheses.

Using structure I as an energy reference, the remaining pathways (A, B, and C) are also examined

(Figure 2-7). Complex II acts as the branching point between aldol-first mechanisms A and B, and

ketene-first mechanism C. Species II can isomerize to III-A (Figure 2-6), and then isomerize again

to 7-membered metallacycle IV-A (TS-III-A). Here paths A and B separate. In path A,

benzaldehyde coordinates to IV-A to create V-A, which then undergoes inner-sphere aldol

addition (TS-V-A) to yield tetracoordinate species VI-A. In path B, a benzaldehyde-BEt3 complex

adds to IV-A (TS-IV-B), to yield V-B.

Page 40: Computational Investigation of Chemical Reactions ...

22

In path C, rather than isomerize to III-A, species II instead undergoes ketene elimination (TS-II-

C) to create ketene complex III-C. Complex III-C then undergoes carbocyclization (TS-III-C),

to yield nickel enolate IV-C. As the formation of IV-C is highly exergonic (-28.5 kcal/mol relative

to I), its formation is expected to be irreversible under the studied conditions. Furthermore, the

relatively low barrier for the formation of IV-C (highest barrier process is TS-II-C, 15.0 kcal/mol)

compared to the barrier for isomerization to a 7-membered metallacycle (TS-III-A, 16.4 kcal/mol),

or the barrier for direct insertion of an aldehyde (TS-II-D, 25.6 kcal/mol), means that the formation

of IV-C in pathway C is expected to outcompete pathways A, B, and D. Based off of the potential

energy surfaces detailed in Figures 2-6, and 2-7, therefore, it was found that ketene-first pathway

C is expected to be the dominant reaction pathway in the studied reaction.

2.4 Determining the Fates of Complexes VI-A, V-B, and IV-C

Figure 2-8. Aldol addition in pathway C. Energies are given in kcal/mol, with enthalpies listed in parentheses.

After carbocyclization to form intermediate IV-C, the aldol addition in pathway C is expected to

occur through an inner-sphere pathway (Figure 2-8). Coordination of benzaldehyde to IV-C yields

Page 41: Computational Investigation of Chemical Reactions ...

23

complex V-C, which can then isomerize to O-bound enolate species VI-C. The O-bound enolate

can then engage in rapid aldol addition (TS-VI-C), to yield aldol adduct VII-C. After aldol

addition, catalyst turnover can then be accomplished by reaction of VII-C with triethylboron.

Figure 2-9. Carbocyclization in paths A and B. Energies are given in kcal/mol, with enthalpies listed in parentheses.

Though Figure 2-7 shows that ketene-first pathway C is expected to be the major pathway in the

three-component coupling of an alkyne, aldehyde, and enoate, the evaluation of pathways A, and

B is still beneficial (Figure 2-9). In particular, evaluating the carbocyclization of paths A and B

distinguishes between those reaction pathways being viable mechanisms that might emerge under

perturbation of the reaction conditions, or being unfeasible mechanisms that should not be

Page 42: Computational Investigation of Chemical Reactions ...

24

considered in the future. Both intermediates VI-A and V-B eventually converge to ethyl-nickel

species VIII-A. In path A, coordination of triethylboron to VI-A yields complex VII-A. This

species can then transfer an ethyl group to the nickel center to create VIII-A. In pathway B,

isomerization to VII-A (TS-V-B) is a higher barrier process compared to direct ethyl transfer (TS-

VI-B) to yield complex VIII-A. After formation of VIII-A, carbocyclization (TS-VIII-A) can

proceed, resulting in carbocycle IX-A, which is expected to be capable of extruding a unit of

phenoxide, and ultimately turning over the nickel catalyst.

While pathway C is expected to be the dominant pathway for catalyst activation, pathway A has

an overall barrier that is only slightly higher in energy. The highest overall barrier for path A is

nickel ethylation (TS-VII-A, Figure 2-9) at 18.5 kcal/mol, 2.5 kcal/mol shy of the largest net

barrier for pathway C prior to carbocyclization (TS-II-C, Figure 2-7). Pathways B and D both

have much larger overall barriers (path B: TS-VI-B, 24.5 kcal/mol, path D: TS-II-D, 25.6

kcal/mol), and can be considered to be much less feasible reactions compared to paths A and C.

2.5 Considering the Reactivity of α-Substituted Enoates

Figure 2-10. Formation of a linear side product in the studied reaction using an α-substituted enoate. The existence

of such a product implicates formation of a 7-membered metallacycle (boxed).

Page 43: Computational Investigation of Chemical Reactions ...

25

The potential energy surfaces shown in Sections 2.3 and 2.4 detail a three-component coupling

reaction involving a β-substituted enoate. However, it is already known that α-substituted enoate

2-3 is capable of producing a linear side product, which implicates the existence of a 7-membered

metallacycle (Figure 2-2). Furthermore, in the development of the three-component coupling

reaction, Aireal Jenkins was also able to identify the formation of a linear product using 2-3 (Figure

2-10). These points of evidence suggest that the mechanism of the reductive [3+2] cycloaddition

between an enoate and an alkyne, and by extension, the three-component coupling reaction

studied, go through different mechanisms, depending on the substitution pattern of the enoate.

Figure 2-11. The phenoxide moiety moves close to the α position (highlighted in gray) after ketene elimination.

Increasing the steric hinderance at that position is expected to destabilize complex III-C.

Closer examination of the geometry of the ketene elimination product (III-C, Figure 2-7) suggests

a rationale for why such a perturbation in mechanism might take place. Figure 2-11 details the

geometry of III-C. Notably, after eliminating the phenoxide moiety to form a ketene, the

phenoxide in the resulting complex is oriented close (2.3 Å) to the hydrogen in the α position

(highlighted in gray). If the α position were to become more sterically crowded, it could be

imagined that ketene elimination would become more difficult.

Page 44: Computational Investigation of Chemical Reactions ...

26

Figure 2-12. The potential energy surface for the first few steps of the [3+2] cycloaddition between an alpha

substituted enoate and an alkyne. Energies are given in kcal/mol, with enthalpies listed in parentheses.

Based off of this observation, the early stages of metallacycle isomerization were investigated

computationally for the three-component coupling reaction involving starting enoate 2-3. The

potential energy surface for these initial steps are shown in Figure 2-12. After isomerization of

metallacycle α-I to rotamer α-II, the complex can either undergo ketene elimination to form

complex α-III-C, or isomerization to yield 7-membered metallacycle α-IV-A. Unlike the reaction

of a β-substituted enolate, metallacycle formation is expected to outcompete ketene elimination

when an α-substituted enoate is used, as both the of the transition states associated with the

formation of the metallacycle (α-TS-II-A, 15.3 kcal/mol and α-TS-III-A, 17.2 kcal/mol) are lower

in energy than the transition state for ketene elimination (α-TS-II-C, 17.8 kcal/mol).

The differences between the potential energy surfaces of the α- and β-substituted enoates can be

more clearly seen if the respective surfaces are lined up next to each other (Figure 2-13). While

Page 45: Computational Investigation of Chemical Reactions ...

27

changing from a β-substituted enoate to an α-substituted one increases the barrier for isomerization

to the 7-membered metallacycle by 0.8 kcal/mol (TS-III-A vs α-TS-III-A), the barrier for ketene

elimination increases by a much larger amount, 2.8 kcal/mol (TS-II-C vs α-TS-II-C). The 2.0

kcal/mol net swing in energy means that ketene elimination changes from being favored by 1.4

kcal/mol to being disfavored by 0.6 kcal/mol, relative to isomerization to the 7-membered

metallacycle.

Figure 2-13. Comparing the barrier to ketene elimination with isomerization to the 7-embered metallacycle for both

the α- and β-substituted enoates.

Even though metallacycle formation is now capable of outcompeting ketene elimination, this does

not guarantee that the three-component coupling of 2-3 goes through an aldol-first type

mechanism. It is possible that ketene elimination could outcompete steps further along the

Page 46: Computational Investigation of Chemical Reactions ...

28

potential energy surface of the aldol-first path, such as triethylboron addition. However, the change

does explain how linear side products that necessitate the formation of a 7-membered intermediate

can occur when 2-3 is used as an enoate, and may also explain why linear side products aren’t

typically seen using β-substituted enoates.55

2.6 Summary and Conclusions

In summary, the mechanism of a three-component coupling reaction (Figure 2-4) was studied.

Prior work from both Montgomery50 and Ogoshi51 has shown that multiple mechanisms are

possible in the reductive [3+2] cycloaddition of an enoate and alkyne, leaving the mechanism of

the three-component coupling reaction developed by Jenkins as ambiguous. In evaluating the

potential energy surface of the three-component coupling, it was found that both an aldol-first

mechanism involving a 7-membered metallacycle, and a ketene-first mechanism involving a

ketene intermediate, are feasible reaction pathways. The ketene-first mechanism (labelled pathway

C) appears to be preferred, due to the ability of ketene elimination to outcompete isomerization to

the 7-membered metallacycle. However, a precarious balance exists between path C and an inner-

sphere aldol-first mechanism (labelled pathway A). Small perturbations in the reaction conditions,

such as using an α-substituted enoate rather than a β-substituted one, appear to be capable of

shifting the mechanism towards aldol-first pathway A. This carries implications for any future

work with nickel-catalyzed reductive [3+2] cycloadditions, including Chapter 3 of this thesis. This

study also showcases the limits of presuming a catalytic mechanism based off of similar prior

work. It is possible for two slightly different substrates to react with the same catalyst to make the

same corresponding product, but to do so through completely different mechanisms.

Experimentalists should use this work as an example for why one should use caution when

applying their untested mechanistic postulations to new reactions.

Page 47: Computational Investigation of Chemical Reactions ...

29

Chapter 3: Elucidation of the Activation Mechanism of Air-Stable

Nickel(0) Catalysts

Some of the material presented in this chapter has been reported in ACS Catalysis.56

3.1 Motivation for the Development of Air-Stable Nickel Pre-Catalysts

Homogeneous nickel catalysis has seen considerable advancements over the past two decades.57–

59 Nickel has played a role in transformations such as C-C bond formation,60–63 C-N bond

formation,64,65 C-O bond formation,65 C-O activation,61,66–68 C-H activation,69–71 decarbonylative

couplings,72 π-component couplings,54,73 and polymerization reactions.74 While related group 10

metal palladium was the dominant catalyst for many of these transformations, resulting in the 2010

Nobel Prize in Chemistry,75 nickel has seen an increase in interest, not only as it is over 5,000

times more abundant in the earth’s crust,76 but also because it is unique suited for reactions such

as alkyl cross-coupling or π-component coupling.57

Despite the great promise of nickel catalysts, there are still many barriers to its adoption. Compared

to palladium, nickel generally requires higher catalyst loadings.77 Additionally, many of the

commonly used nickel precursors, such as Ni(COD)2 (bis(cyclooctadiene)nickel(0)), are air-

sensitive, and require the use of a glovebox.57 Both of these problems can potentially be addressed

through the development of new, air-stable nickel pre-catalysts.

The use of pre-catalyst, with the desired ligands already bound to the metal center, is an effective

strategy for lowering the amount of catalyst needed for a chemical transformation.77 In-situ

complex formation introduces ambiguity about the nature and quantity of active catalyst involved

Page 48: Computational Investigation of Chemical Reactions ...

30

in the reaction. A portion of the metal introduced may fail to ligate, or may be ligated in an

unproductive manner. If only a small portion of the metal and ligand introduced form an active

catalytic complex, then observed catalyst turnover number will be artificially lowered. The failure

of the majority of introduced metal to form a competent catalyst artificially lowers the observed

turnover number. Consequently, by using a pre-catalyst, where all of the material is already ligated,

similar yields to in-situ procedures can be obtained using much lower catalyst loadings, or much

higher yields can be obtained with the same catalyst loading.

Figure 3-1. Use of a pre-catalyst dramatically improves the yield of 3-1. Crystal structure taken from reference 78.

An example of this effect is illustrated by a publication from the Montgomery group.78 A nickel(0)

catalyst in conjunction with small NHC (N-heterocyclic carbene) ligand ITol was found to be an

effective combination for the deoxygenative coupling of an unsaturated aldehyde or ketene with

an alkyne, to form skipped diene species 3-1 (Figure 3-1). While the in-situ formation of the

catalyst gave poor yields, use of pre-ligated complex 3-2 gave much improved yields. The success

of 3-2 as a catalyst in this transformation demonstrates the power of using a pre-catalyst: the same

Page 49: Computational Investigation of Chemical Reactions ...

31

amount of metal and ligand are used, but using an isolable, pre-ligated species ensures that all of

the material added is catalytically active.

Figure 3-2. (Top) Air-stable Ni(II) catalysts require a transmetallation reagent. (Middle) Tradeoffs in the development

of Ni(0) pre-catalysts. (Bottom) Molecular orbital diagram for the π backbonding interaction.

Many of the reactions developed by the Montgomery group have utilized a low-valent nickel

complex in conjunction with an NHC ligand, largely for the purposes of coupling π-components.54

While air-stable pre-catalysts that include a ligated NHC ligand already exist, most of them are

Ni(II) species such as 3-379 and 3-480. Such Ni(II) precursors are competent for reactions that

involve a transmetallating reagent, and do not have a clear means of activation for reactions that

Page 50: Computational Investigation of Chemical Reactions ...

32

only involve π-components and silanes (Figure 3-2, Top). For this reason, Ni(0) pre-catalysts

bound to π-acceptor ligands such as 3-5,81 3-6,82 3-2,78 or 3-783 were considered. In the choice of

olefin, a tradeoff emerges between catalyst activity and air-stability (Figure 3-2, Middle). The

more electron-poor the ligand, the better it is able to engage in a π-backbonding interaction, where

the d orbitals of the metal interact with the π* antibonding orbitals of the olefin (Figure 3-2,

Bottom). A greater the backbonding interaction increases the stability of the complex, as it pulls

electron density away from the metal center, reducing the metal’s susceptibility towards oxidation.

However, an increased backbonding interaction will also increase the π-acceptor’s binding energy,

make fumarate dissociation more difficult, and preventing the generation of an active catalyst.

3.2 Investigating a Diverse Set of Nickel Fumarate Complexes

In this context, the development of an air-stable nickel catalyst with a carbene ligand already bound

was sought. Fumarate complexes such as 3-7 were chosen as a starting point in this investigation,

as such complexes were reported to be air-stable,83 but unreactive towards π-component couplings

in our hands. Given the trends seen with olefin ligands, it was expected that an inverse relationship

between reactivity and stability would occur. With this hypothesis in hand, the design strategy was

to develop fumarate catalysts that might have greater steric clashes with the NHC ligand,

weakening the fumarate-Ni interaction enough to enable ligand exchange (Figure 3-3, Top).

Visiting student Santiago Cañellas and Montgomery student Alex Nett synthesized nine different

fumarate complexes using the IMes NHC, and then they tested their ability to engage in the

reductive coupling reaction to produce 3-9 (Figure 3-3, Bottom). In general, all of the tested

fumarate complexes were found to be more active than Cavell’s originally reported complex (I).83

Most of the aryl fumarate complexes gave low to moderate yields (E, F, G, H), but complexes of

fumarates A and C gave high yields of 3-9, as well as complexes of alkyl fumarates B and D.

Page 51: Computational Investigation of Chemical Reactions ...

33

These 4 fumarate complexes were then subsequently tested for their air stability, and all 4 (A

through D) were found to be air-stable.

Figure 3-3. (Top) Different fumarate complexes were synthesized. (Bottom) The different complexes were tested for

their ability to catalyze a reductive coupling reaction.

However, the discovery of air-stable, reactive Ni(0) IMes fumarate catalysts was not found to be

easily transferred to other NHCs. For instance, Montgomery group student Amie Frank was unable

to use complexes of fumarate A with a chiral NHC,84 or smaller carbene ligand BAC85 for a

reductive coupling reaction, even at higher catalyst loadings compared to complex 3-8-A. The

incompetence of catalysts 3-10-A and 3-11-A suggested that a greater understanding of the catalyst

activation process was necessary, and that the identity of the fumarate used may need to be tuned

based on the NHC used in the complex.

Page 52: Computational Investigation of Chemical Reactions ...

34

Figure 3-4. Other complexes of fumarate A are not competent in reductive couplings.

3.3 Overturning the Original Hypothesis that Catalyst Activation Occurs via Dissociation

Figure 3-5. The mechanism of the reductive coupling reaction. Originally ligand dissociation was believed to be the

means of catalyst activation.

In the synthesis of complexes 3-8, the design strategy was to change the identity of the R group to

increase steric repulsion between the fumarate and metal center. Given the mechanism of the

Page 53: Computational Investigation of Chemical Reactions ...

35

reductive coupling reaction (Figure 3-5), it was thought that the metal center needed to have 2 free

coordination sites in order to perform catalysis. This necessitates some form of ligand dissociation

to allow for binding of the alkyne and aldehyde starting material.

Figure 3-6. The relationship between the free energy of ligand exchange and the yield of 3-9 in reductive coupling

attempts with the corresponding catalyst. Computational details are included in the appendix.

Notably, the hypothesis that dissociative ability of a fumarate complex determines its catalytic

ability can be easily tested computationally. The free energy of ligand exchange was calculated

for each fumarate complex listed in Figure 3-3, and then that energy was compared to the yield of

3-9 that was observed experimentally. If it was the case that ligand dissociation determined

catalytic ability, it would be expected that some form of statistical relationship should emerge

Page 54: Computational Investigation of Chemical Reactions ...

36

between the free energy of dissociation and the observed yields. However, no relationship between

reaction yield and computed binding affinity was found (Figure 3-6).

The lack of correlation seen in Figure 3-6, suggested that a fumarate dissociation mechanism was

not the mechanism of catalyst activation, as was originally thought. As fumarate dissociation for

all of the studied complexes 3-8 was found to be endergonic, it can be concluded that all of the

studied fumarates can act as a catalyst poison. Even if the weakest bound fumarate, D, is

considered, it would be expected that dissociation of 3-8-D to form a nickel-aldehyde-alkyne

complex would raise the energy of the subsequent oxidative addition step by 13.8 kcal/mol,

slowing the reaction down by a factor of over 10 billion. It follows then, that since many members

of catalysts 3-8 are capable of producing 3-9, it must be the case that their corresponding fumarates

are consumed prior to formation of product 3-9. Were this hypothesis to be true, it should be

possible to isolate the products of such a fumarate consumption reaction. Using this prediction,

Montgomery student Ellen Butler investigated the reactivity of catalyst 3-8-A in greater detail.

Under conditions similar to the reductive coupling reaction, products 3-12 and 3-13 were isolated

by reacting 3-8-A with phenyl propyne, 4-fluorobenzaldehyde, and triethyl silane. (Figure 3-7).

As 3-12 is very similar to the products formed in the three-component coupling studied in Chapter

2, we hypothesized that catalyzed activation would occur through an analogous reaction.

Figure 3-7. The isolation of products 3-12 and 3-13 from catalyst 3-8-A.

Page 55: Computational Investigation of Chemical Reactions ...

37

3.4 Investigating the Activation Sequence of BAC and IMes Fumarate Complexes

Figure 3-8. Possible activation mechanisms for complexes of fumarate A.

In Chapter 2, the formation of 3-component products such as 3-12 was found to occur through

either an aldol-first or ketene-first type mechanism, depending on the identity of the π-component

used in the transformation (Figure 3-8). To distinguish between the mechanisms of catalyst

activation in 3-8-A, as well as to investigate why catalyst 3-10-A fails to successfully activate, the

mechanism of catalyst activation of both the BAC and IMes complexes with fumarate A were

investigated computationally.

Page 56: Computational Investigation of Chemical Reactions ...

38

Figure 3-9. The initial steps of activation for IMes complex 3-8-A (blue and turquoise) and BAC complex 3-10-A (red

and pink). Dark colors (red, blue and black) represent aldol-first path A. Light colors (pink, turquoise, and gray)

represent ketene-first path B. Energies are given in kcal/mol, with enthalpies listed in parentheses. Computational

details are included in the appendix.

The potential energy surfaces of the early steps of aldol-first path A and ketene-first path B for

both IMes catalyst 3-8-A (pathways shown in blue and turquoise, labelled as IMes) and BAC

catalyst 3-10-A (pathways shown in red and pink, labelled as BAC) are shown in Figure 3-9. In

Chapter 2, it was demonstrated that the selectivity between an aldol-first reaction and a ketene-

first reaction is dependent on the barrier of ketene elimination compared to the highest barrier step

in the aldol-first process. The early steps listed in Figure 3-9, therefore, play a large role in

determining the path-selectivity of catalyst activation in 3-8-A and 3-10-A.

Page 57: Computational Investigation of Chemical Reactions ...

39

In the initial steps of path A (dark colors), 5 membered metallacycle I rotates to isomer II, and

then isomerizes to ξ-3 bound III-A (TS-II-A). Complex III-A then isomerizes again (TS-III-A)

to 7-membered metallacycle IV-A. Alternatively, in path B (light colors), isomer II extrudes a unit

of aryloxide (TS-II-B), to create ketene complex III-B. The ketene species can then cyclize (TS-

III-B) to carbocyclic species IV-B.

BAC catalyst 3-10-A and IMes catalyst 3-8-A differ significantly in these early steps. For the BAC

complex, ketene elimination (BAC-TS-II-B, 13.4 kcal/mol) is fast enough that it outcompetes

isomerization of ξ-3 bound BAC-III-A to 7-membered metallacycle (BAC-TS-III-A, 19.5

kcal/mol). After forming ketene complex BAC-III-B, an irreversible carbocyclization can occur

(BAC-TS-III-B, 7.1 kcal/mol), yielding carbocycle BAC-IV-B. Taken together, the larger barrier

height of BAC-TS-III-A (19.5 kcal/mol) compared to BAC-TS-II-B (13.4 kcal/mol) indicates

that BAC complex 3-10-A prefers to undergo catalyst activation through path B.

However, IMes catalyst 3-8-A behaves differently. For the IMes catalyst, isomerization of 5-

membered IMes-II-A to ξ-3 bound IMes-III-A is the slowest step towards the formation of 7-

membered IMes-IV-A (IMes-TS-II-A, 15.1 kcal/mol). Unlike the BAC complex, ketene

elimination (IMes-TS-II-B, 22.0 kcal/mol) is too slow to outcompete isomerization to the 7-

membered metallacycle. Due to the higher barrier for ketene formation, and the lower barriers for

isomerization to the 7-membered metallacycle, path A is still a possible activation mechanism for

the IMes catalyst, and would be expected to be the preferred mechanism of activation if it forms

an intermediate that is sufficiently exergonic compared to starting metallacyle IMes-I.

As the preliminary steps for paths A and B suggested that IMes catalyst 3-8-A and BAC catalyst

3-10-A undergo activation by different mechanisms, we hypothesized that the difference in

mechanism can explain why 3-8-A is a competent catalyst, but 3-10-A is not. In order to evaluate

Page 58: Computational Investigation of Chemical Reactions ...

40

this hypothesis, the progress of path A with 3-8-A and path B with 3-10-A was traced further down

their respective potential energy surfaces.

Figure 3-10. Pathway A of the activation of IMes catalyst 3-8-A, part 1. Energies are given in kcal/mol, with

enthalpies listed in parentheses. Computational details are included in the appendix.

In the case of catalyst 3-8-A, path A provides a means to release a potential active catalyst. Figures

3-10 and 3-11 detail the pathway for catalyst release. Seven-membered metallacycle IMes-IV-A

can ligate to an aldehyde (IMes-V-A, Figure 3-10), and can then undergo an aldol reaction (IMes-

TS-V-A) to yield complex IMes-VI-A. Notably, this process has a slightly higher barrier than the

isomerization process (15.9 kcal/mol), but still outcompetes ketene elimination (IMes-TS-II-B,

22.0 kcal/mol, Figure 3-9). After aldol addition, ligation of tetrahydrofuran to complex IMes-VI-

Page 59: Computational Investigation of Chemical Reactions ...

41

A is possible (IMes-VI-A-THF), but an irreversible hydrosilylation (IMes-TS-VI-A) can occur,

yielding complex IMes-VII-A. Complex IMes-VII-A can then rearrange to nickel hydride species

IMes-VIII-A. Subsequently, IMes-VIII-A could either undergo ester reduction (IMes-TS-VIII-

Z, 7.0 kcal/mol, Figure 3-11) to form IMes-IX-Z, or undergo a carbocyclization event (IMes-TS-

VIII-A, -7.8 kcal/mol). The latter process is preferred by a significant margin, and leads to nickel

alkoxide species IMes-IX-A, which can easily extrude an alkoxide to yield compound 3-13 and

an activated catalyst.

Figure 3-11. Pathway A of the activation of IMes catalyst 3-8-A, part 2. Path Z details the possible ester reduction

reaction, and is shown in seafoam green. Energies are given in kcal/mol, with enthalpies listed in parentheses.

Computational details are included in the appendix.

While IMes catalyst 3-8-A has been shown experimentally to be a competent catalyst in the

production of 3-9 (Figure 3-3), the same cannot be said for BAC catalyst 3-10-A (Figure 3-4).

Computational investigation of path B of catalyst 3-10-A reveals a putative reason why that might

Page 60: Computational Investigation of Chemical Reactions ...

42

be the case (Figure 3-12). Cyclization of ketene complex BAC-III-B yields carbocycle BAC-IV-

B (Figure 3-9). Notably, the presence of a proximal ester moiety in BAC-IV-B allows for direct

coordination of the ester to the nickel center (BAC-V-B, Figure 3-12). In effect, by occupying a

coordination site, the proximal ester prevents the coordination of an aldehyde that is reported in

Chapter 2. The stability of BAC-V-B, and the additional chelation that occurs in the complex, can

explain why BAC catalyst 3-10-A is ineffective.

Figure 3-12. (Left) Complex BAC-IV-B can rearrange into carbonyl bound complex BAC-V-B. (Right) A three-

dimensional representation of complex BAC-V-B.

3.5 Preliminary Results Towards the Development of an Air-Stable Nickel(0) BAC Catalyst

The difference in activation mechanism between BAC complex 3-10-A and IMes complex 3-8-A

can explain why 3-8-A is active in reductive coupling reactions, but 3-10-A is not. With a

relationship between catalyst activation mechanism and catalyst efficacy suggested, the next

logical step is to develop a catalyst that goes through an “aldol-first” mechanism, so as to avoid

catalyst trapping by developing a species such as BAC-V-B (Figure 3-12). As the fate of catalyst

3-10-A is determined in part by the barrier of ketene elimination (BAC-TS-II-B, Figure 3-9)

compared to isomerization (BAC-TS-III-A, Figure 3-9), the transition states for ketene

Page 61: Computational Investigation of Chemical Reactions ...

43

elimination and metallacycle isomerization were evaluated for a variety of different BAC fumarate

complexes (Figure 3-13). Of the 12 fumarates tested, only 3 (B, I, J) were found to favor

metallacycle isomerization over ketene elimination. Additionally, 3 more fumarates (C, K, M) had

metallacycle isomerization within 1 kcal/mol of ketene elimination. In all 6 cases, the difference

in energy between the two possible isomerization mechanisms is within the reported error for the

functional used in the analysis, ωB97X.13 For this reason, the 6 fumarates listed (B, C, I, J, K, M)

are viable candidates for experimental study.

Figure 3-13. Preliminary efforts towards developing a BAC fumarate catalyst. Fumarates that favor metallacycle

isomerization over ketene elimination are shown in blue. Fumarates where metallacycle isomerization is within 1

kcal/mol or less than ketene elimination are shown in purple. Listed energies are free energies in kcal/mol, and are

relative to BAC-I (Figure 3-9). Computational details are included in the appendix.

The fumarate R groups where metallacycle isomerization is competitive with ketene elimination

have a diverse steric and electronic profile, ranging from small substrates (I, J) to bulky substrates

(K, M), and including both aryl (C, M) and alkyl (B, I, J, K) groups. A clearer pattern exists for

the fumarates that clearly favor ketene elimination (D, E, F, G, H, L). Almost all of these fumarate

Page 62: Computational Investigation of Chemical Reactions ...

44

R groups are aryl species, and are generally non-bulk aryl groups. For instance, in all of the tested

aryl substrates that lacked an ortho substituent (E, G, H), ketene elimination is favored over

metallacycle isomerization by at least 6 kcal/mol.

Additionally, the competitiveness of metallacycle formation appears to be driven more by

destabilizing the ketene elimination step, rather than stabilizing the isomerization process. The

lowest barrier ketene elimination step (10.6 kcal/mol, F) is much lower in energy than the lowest

barrier isomerization step (17.8 kcal/mol, M). This presents a problem in catalyst development, as

if metallacycle isomerization needs to be fast enough to enable catalyst release. Interestingly, all

of the aryl fumarates tested have lower barriers for metallacycle isomerization than any of the alkyl

fumarates tested.

While the data presented here represent only a preliminary study into the factors that affect

selectivity for path A over path B, some key insights into the factors that affect path selectivity can

be gleaned. Alkyl groups appear to generally favor isomerization over elimination, but the barriers

for both processes appear to be higher compared to aryl groups. Aryl groups, in turn, appear to

generally favor elimination over isomerization, but both processes are lower in energy. Steric bulk

also appears to have a positive effect in making the isomerization process more competitive than

elimination, which is consistent with what was observed in Chapter 2 with α-substituted enoates.

Taken together, the most promising fumarate tested in Figure 3-13 appears to be fumarate M. This

fumarate has the lowest barrier for metallacycle isomerization among the tested set, and while

ketene elimination is favored, it is only favored by an amount within the error of the calculation

(0.6 kcal/mol). Future developments in creating a bulkier version of M may allow for the

development of an air-stable Ni(0) BAC complex.

Page 63: Computational Investigation of Chemical Reactions ...

45

3.6 Summary, Conclusions, and Outlook

In summary, the mechanism of activation of nickel NHC fumarate complexes has been identified.

The original hypothesis of ligand displacement was disproven by comparing the binding energies

of the fumarate to the yield of product observed experimentally. Based on these initial

computational results, a new hypothesis of fumarate consumption during catalyst activation was

crafted, ultimately leading to the experimentalists identifying products of the associated fumarate

consumption reaction. The isolation of fumarate consumption species 3-12 and 3-13 determined

that catalyst activation occurs through a reaction analogous to the three-component coupling

reaction detailed in Chapter 2. Using the three-component coupling as a guide, catalyst activation

through both an aldol-first and ketene-first pathway was examined computationally. It was found

that active IMes catalyst 3-8-A undergoes an aldol-first catalyst activation mechanism, but inactive

BAC catalyst 3-10-A undergoes a ketene-first activation mechanism, where it becomes trapped as

ester-bound complex BAC-V-B. Preliminary work has been done to find a fumarate that can be

paired with BAC that would ensure that the BAC catalyst is activated through an aldol-first

mechanism, and avoid catalyst trapping.

Page 64: Computational Investigation of Chemical Reactions ...

46

Chapter 4: Synergistic Effects Between Radical Attack and Applied

Force in the Depolymerization of Poly(Acrylic Acid)

4.1 The Potential Energy Surface Under Applied Force

Chemists have used tools such as photoexcitation, applied electric potentials, or simply the

manipulation of temperature to alter the kinetics and thermodynamics of chemical reactions. The

application of an external mechanical force represents another avenue to introduce energy into a

chemical system.86 Through the introduction of a force bias, chemical reactions that are

endergonic, or are too high in barrier to proceed at a reasonable rate, become accessible to

chemists.

Figure 4-1. The additive effect of applied force on the Morse potential. The distortion due to applied force is greater

at larger C-C bond distances.

Bond scission events are a classic example of how an applied mechanical force can change the

nature of chemical reaction. A carbon-carbon (C-C) bond has a homolytic bond dissociation

energy of approximately 90 kcal/mol,87 and due to this high endothermicity, C-C bond scission is

Page 65: Computational Investigation of Chemical Reactions ...

47

not expected to occur spontaneously. However, if an external force is applied, the energy of a

chemical system decreases as the two carbons move further and further apart. This means that

eventually, at a large enough distance, C-C bond scission becomes exothermic, and

thermodynamically favorable. Figure 4-1 details the effect that applied force has on a potential

energy surface. If we consider the C-C bond as adopting a Morse potential (A), we can then

combine that potential with an external force (B) to create a force-modified potential (C). In

potential C, bond scission is exothermic, but there a barrier associated with the bond scission event,

that is dependent on the amount of force required.

Figure 4-2. Estimating the tensile of a C-C bond using the Morse method.

Considering a chemical bond as a Morse oscillator has the additional advantage of allowing for an

estimation of the tensile strength of the bond (FT, the force required to break the bond). The first

derivative of the Morse potential is the force that the potential exerts on the chemical bond (internal

Page 66: Computational Investigation of Chemical Reactions ...

48

force). Applying an external force, then, can be seen as counteracting the internal forces of the

bond. If it is assumed that a bond breaks when the no barrier for scission exists, it can then be

inferred that the tensile strength of a bond is simply the maximum of the first derivative of the

Morse potential. Using the functional form of the Morse potential, the tensile strength can be

derived, yielding a dependence only on the bond dissociation energy (E0 ), and bond force constant

(𝑘) (Figure 4-2).88

Figure 4-3. The expected effects of force according to the tilted potential energy surface model.

Of course, in reality, bond scission events can occur even if a kinetic barrier exists. Preliminary

work by Zhurkov has shown that applied force has a linear effect on such a kinetic barrier, enabling

a low barrier for reaction (and thus faster rate of reaction) with an increase in applied force.89 Force

is also expected to change the geometries of the intermediates involved. This is adequately

described by the tilted potential energy surface model (TPES) (Figure 4-3). In general, with

increasing external force, it is expected that the starting geometry (Start) of the material becomes

more product-like with increasing force, and the transition state (TS) becomes more reactant-like.

Page 67: Computational Investigation of Chemical Reactions ...

49

The net effect of this process is that the starting structure and the transition state become more

similar with increasing applied force.

Computational methods offer an avenue to investigate mechanochemical processes.90,91 A

common method of computational investigation is the COGEF (constrained geometries simulate

external force) method. In COGEF, the potential energy surface of stretching a molecule is mapped

using a series of geometry optimizations with an increasing distance constraint. In the seminal

paper of the method, Beyer uses the information derived from this potential map to estimate the

tensile strengths of various bonds, including C-C, C-N, and Si-O at various timescales.92 Notably,

the method cannot identify the transition state of the bond scission event, as it infers activation

energies from a theoretical force-modified potential that is based off of the curvature of the

potential energy surface under no force (similar to the Morse model described in Figure 4-2).

Despite this limitation, COGEF still finds use in exploring mechanochemical processes; for

instance, in comparing the COGEF potential of unprotonated vs protonated dimethyl ether, Beyer

finds that proton affinity of an ether increases with applied force,93 supporting a previous Carr-

Parinello dynamics study that demonstrated the role of aqueous solvation in the depolymerization

of poly(ethylene glycol).94 In complicated systems such as lignin, ab initio steered molecular

dynamics (AISMD) have been used to identify the bonds most susceptible to scission during

depolymerization.90 AISMD has also been used to identify the heterolytic character of poly(o-

phthalaldehyde) depolymerization.95

While popular tools such as COGEF are unable to provide a transition-state geometry,92 recent

developments in the growing string method allow for the identification of transition states in the

presence of an applied force (F-GSM).96 In this chapter, F-GSM is the method of choice for

interrogating bond scission events, allowing for detailed analysis on the transition state geometries

Page 68: Computational Investigation of Chemical Reactions ...

50

of the processes. The tool works in a similar fashion to GSM (see Section 1.5), with the addition

that the energy of a given state is modified by the term −𝐹 ∗ ∆𝑥, where 𝐹 is the applied force, and

∆𝑥 is the change in distance between the atoms upon which force is applied.

4.2 Plastic Recycling Using Ultrasonic Depolymerization

Plastic waste accumulation in the environment occurs on a massive scale, where it is predicted that

the mass of plastics in the ocean will exceed the total biomass of fish by the year 2050.97 This

problem might be somewhat alleviated by recycling, but the United States only recycles roughly

9.1% of plastic waste, a low figure compared to recycling rates for paper (66.6%), glass (26.4%),

and metals (34.3%).98 The recycling rate of plastics is limited99 by methods that mechanically

repurpose certain plastics for new applications, usually resulting in lower quality materials which

reduces economic incentives to recycle. For this reason, chemical methods of recycling such as

depolymerization attract interest due to their potential to reach a greater scope of materials and

create more valuable recycled products. By developing and expanding recycling technologies such

as depolymerization, the recycling rate for plastics could be greatly improved, ultimately having a

great positive impact on the environment.

A promising depolymerization strategy is to use the mechanochemical technique of ultrasonic

irradiation. (Figure 4-4, A).100 In ultrasonic depolymerization, an acoustic field with frequencies

of greater than ~20 kHz is applied to a solution of polymer.88,91,101–105 The pressure variations

imparted by the high frequency sound waves form cavitation bubbles in solution that ultimately

collapse (Figure 4-4, B), producing a shear force that is capable of tearing polymer chains apart

(Figure 4-4, C). Additionally, pyrolytic reactions in the cavitation bubbles are known to produce

free radicals (Figure 4-4, D). These radicals are believed to accelerate the process of polymer

Page 69: Computational Investigation of Chemical Reactions ...

51

degradation,106 as introduction of an external radical source has been shown to increase the

breakdown rate.107–114

Figure 4-4. Recycling of polymers is enabled by ultrasonic depolymerization.

Prior studies have considered the roles of mechanical force and radical species—each separately—

in the process of ultrasonic irradiation. Based on studies with degassed solvent,115 and the observed

propensity for midpoint scission to occur,116 the mechanical forces generated though cavitation

bubble collapse are thought to be the dominant source of bond scission.88,90,101 Mechanical scission

generally occurs through a homolytic pathway,88,101,102 resulting in the formation of two free

macroradicals. These homolytic cleavage events have been observed by monitoring the

Page 70: Computational Investigation of Chemical Reactions ...

52

stoichiometric consumption of a radical trap.117 Macroradicals have also been directly observed

through EPR studies during the sonication process.118

As a consequence of the depolymerization mechanism largely being mechanical scission,

depolymerization is expected to occur until the molecular weight of the polymer converges to a

limiting value. The origin of this effect is best explained by a model developed by Okkuama and

Hirose.119 In their model, the force that a polymer experiences is due to the friction of the monomer

units with the surrounding solvent. Consequently, the overall force that a given polymer strand

feels is proportional to the number of monomers in a given strand. While the force experienced by

each polymer strand shrinks as the strand gets shorter, the tensile strength of the polymer remains

constant. The limiting length of a given polymer, then, is simply the length at which the forces

experienced by the polymer strand are insufficient to cause further chain scission. Simon has used

this model to accurately simulate the time evolution of polymer degradation.120

Separate from the formation of polymer macroradicals during ultrasonic polymerization, small

radical species are also known to form during the cavitation process.121 As cavitation bubbles

grow, the frequency of their vibration increases, heating the interior of the bubble to several

thousand degrees Kelvin. At such high temperatures, volatile compounds in the bubble, such as

solvent molecules, can pyrolytically disproportionate to small radicals. These radicals are known

to escape from the cavitation bubble and enter solution, where they can react with dissolved

materials.122–125 The free radicals that are generated are also known to initiate radical

polymerization in solutions of monomer.126

Quenching studies have shown that free radicals can play a role in the depolymerization process

as well. Koda106 has shown that the addition of radical scavenger tert-butanol inhibits the ultrasonic

degradation of various polymers. Likewise, the introduction of radical-generating species is known

Page 71: Computational Investigation of Chemical Reactions ...

53

to accelerate ultrasonic degradation processes. The Chen group, as well as others, have reported

depolymerization processes that combine Fenton chemistry with ultrasound.107–111 Yao has

reported a synergistic effect of bubbling ozone during the ultrasonic irradiation of chitosan

solutions.113 Gogate has shown that addition of oxidant potassium persulfate is beneficial in the

degradation of guar gum using hydrodynamic cavitation.112 This group also found that addition of

H2O2 or ozone aids in the ultrasonic depolymerization of poly(acrylic acid).114

The interplay between radical and ultrasonic degradation plays an important role in breaking down

polymers, but the specific mechanism(s) involved in this synergy is not fully explicated. Koda,106

for instance, proposes random chain scission events as an explanation for why free radicals affect

ultrasonic polymer degradation. In Koda’s report, however, it was also observed that the limiting

molecular weights of the degraded polymers change in the presence or absence of a radical

scavenger. The implication of this observation is that in the presence of radicals, there is a change

in how the polymer responds to mechanical force, as the limiting length effect is due exclusively

to mechanochemical degradation. Based on this implication, it appears that the radical species

formed during ultrasonic degradation play a role beyond that of enabling simple random chain

scission events.

Figure 4-5. This chapter explores the synergistic interplay between radical attack and tensile force.

This chapter seeks to provide a clear, atomistic picture of how radical species affect the

sonochemical depolymerization process (Figure 4-5). The synergistic effect of radical attack and

Page 72: Computational Investigation of Chemical Reactions ...

54

tensile force in the ultrasonic depolymerization of poly(acrylic acid) (PAA) will be described. PAA

was chosen as it is broadly used as a super-absorbent polymer, as well as in paints, dentistry, and

other applications.114 The origin of the change in limiting length observed during depolymerization

can be explained though the interplay of radical attack and application of mechanical force. Radical

attack will be shown to have a weakening effect on the tensile strength of the polymer, which will

affect the limiting length. The effects of force on the transition state energy of bond scission will

be quantified, allowing for the inclusion of thermal effects. Finally, it will be shown that radical

attack causes the polymer to respond to force in a manner distinct from what is normally seen in

mechanochemistry.

4.3 Evaluating the Impact of the Weak Bond Effect Using a Morse Potential

Incorporation of a weak bond into a polymer backbone is expected to accelerate the rate of chain

scission.91 For example, when Encina and coworkers added peroxide linkages in

poly(vinylpyrrolidone), the rate of ultrasonic degradation increased by a factor of 10.127 Using

diazo-linked polymers, Moore and coworkers provided evidence that cleavage is mostly localized

at weak bonds.128 This “weak bond effect” suggests that radical activation of PAA might synergize

with tensile force to break down this polymer. While the homolytic bond dissociation energy of a

carbon-carbon bond is generally in the vicinity of 90 kcal/mol,87 the heat of (radical)

polymerization for polyolefins is much lower, in the range of 10-30 kcal/mol.129 Acrylate

polymerizations are in the vicinity of 19 kcal/mol,129 suggesting that lower forces will break the

weakened C-C bond in activated PAA.

Page 73: Computational Investigation of Chemical Reactions ...

55

Figure 4-6. Model systems used in this study: Degradation of PAA through force alone (PATH-1), and force in

conjunction with radical attack (PATH-2).

Two model systems of PAA were used to quantify the bond-weakening effect of radicals (Figure

4-6). The first is a tetramer of AA with inactivated C-C bonds (1), which can be fragmented via

tensile force into a biradical (3+4). This degradation pathway is called PATH-1. The second (2)

is similar to 1, except one hydrogen atom from the backbone has been removed, for example by

radical abstraction via a sonication-generated hydroxyl. C-C bond breaking for the radical-

activated species is denoted PATH-2.

Figure 4-7. Flowchart for evaluating tensile strength using the Morse method. The complete procedure is described

in the appendix.

These pathways were first examined by treating the polymer scission as a purely mechanical event

using the Morse method. A flowchart detailing this process is shown in Figure 4-7. A library of

conformations of 1 and 2 were first generated. Conformers of 1 were generated using the Confab

Page 74: Computational Investigation of Chemical Reactions ...

56

tool130 in Open Babel,131 which resulted in 8 unique conformers. Conformers for 2 were generated

by abstracting an α-hydrogen from each of the conformers of 1. As hydrogen bonding is known to

affect the heat of polymerization of protic polymers such as PAA,132 and to ensure that the

conformers used in this study are consistent with aqueous solvation, explicit water solvation was

modeled. For each conformer of 1 and 2, the system was surrounded by a 6 Å water shell

containing 260 water molecules. Molecular dynamics (MD) simulations were then used to create

a series of snapshots of the tetramer in an aqueous environment.

Figure 4-8. Example three-dimensional representations of the tetramer (Right) and dimer+dimer (Left) systems used

for this study. These represent two of many snapshots used for this study.

The geometry of selected snapshots was then re-optimized using a non-periodic quantum

mechanical/molecular mechanical (QM/MM) method. After optimizing the geometry of the

snapshots (Figure 4-8, Left) the scissile bond (shown in red, Figure 4-7) was cleaved by applying

a stretching force of 10 nN between the two atoms of the scissile bond using the EFEI (external

force explicitly included) method.133 After bond scission, subsequent re-optimization yielded two

Page 75: Computational Investigation of Chemical Reactions ...

57

PAA dimers (referred to as dimer+dimer snapshots, seen in Figure 4-8, Right). A complete

description of the entire procedure can be found in the appendix.

As denoted in Figure 4-2, the tensile strength of a bond can be determined from its bond

dissociation energy (E0 ) and the bond’s force constant (k ). The bond dissociation energy was

determined by comparing the total energy of each dimer+dimer snapshot to the tetramer from

which it came. By averaging the difference in energy for each set of snapshots, a value for the

enthalpy of bond scission is obtained. To determine the force constant of the scissile bond, a

vibrational constant was determined from each optimized tetramer absent any explicit water using

partial hessian vibrational analysis.134 The calculated force constant values were averaged to yield

the force constant of the scissile bond.

Figure 4-9. The effect of hydrogen atom abstraction on the tensile strength poly(acrylic acid). Computational details

are included in the appendix.

These values were used in the final equation for tensile strength (FT ) (Figure 4-2). The tensile

strengths of 1 and 2 were calculated as 6.6 nN and 2.6 nN, respectively (Figure 4-9), both within

the range of forces available for sonication.88 Hydrogen atom abstraction, therefore, is calculated

Page 76: Computational Investigation of Chemical Reactions ...

58

to reduce the tensile strength of PAA by approximately 4 nN. As the limiting length of polymer

during sonication is proportional to the square root of the polymer’s tensile strength,119 the results

in Figure 4-9 predict that the limiting length of a polymer weakened by radicals should be ~37%

lower than a polymer free of radical defects. These results are on the order of the changes in

limiting length observed by Koda upon suppression of radical formation during the sonication of

poly(ethylene oxide) and polysaccharides.106

The predicted tensile strength of 1 of 6.6 nN using the simple Morse method is outside the range

of modern estimates of C-C bond strength, such as the thermally activation barrier to scission (5-

6 nN),135 or COGEF (4.5-5 nN) models. This discrepancy likely stems from the assumption that

scission only occurs in the absence of a thermal barrier.92 In reality, it is expected that bond scission

would occur once a thermally accessible barrier is attained. Analysis of the bond scission reaction

using F-GSM will identify the nature of this barrier, and allow for a more accurate determination

of the effect of radical abstraction on the transition state of bond scission and tensile strength.

4.4 Determining the Effect of Radical Abstraction on the Transition State of Bond Scission

Polymer degradation by sonication is better described by treating cleavage events as a chemical

reaction. Using this model, the reactant and transition state geometries—as well as the activation

barrier—depend on the amount of force applied to the polymer chain. Using the sets of tetramers

from the previous section, the transition states of bond scission for PATH-1 and PATH-2 were

found using the force-biased growing string method (F-GSM, see appendix for the full process)

Page 77: Computational Investigation of Chemical Reactions ...

59

for a range of applied tensile forces.96 The activation enthalpies as a function of force from F-GSM

are given in Figure 4-10.

Figure 4-10. The relationship between enthalpy of activation and applied force on the polymer chain. Computational

details are included in the appendix.

Given that hydrogen atom abstraction by a hydroxyl radical is fast (ΔH‡ of ~3 kcal/mol) and

irreversible,136 the selectivity between PATH-1 and PATH-2 is expected to be largely determined

by the availability of hydroxyl radicals in solution. However, the rate of bond scission for PATH-

1 and PATH-2 will still be relevant if hydrogen atom abstraction occurs, as the relative rate

between PATH-1 and PATH-2 will determine whether scission will occur at the point of hydrogen

atom abstraction. Essentially, hydrogen atom abstraction can only possibly be beneficial if it

provides a lower barrier pathway for bond scission to occur.

At low levels of applied force, it is expected that the barrier for bond scission in PATH-1 would

be insurmountably high. For this reason, only the activation energies for forces in the range of 3

to 5 nN are shown for PATH-1. A negative linear relationship between the enthalpy of TS-1 (ΔH‡)

Page 78: Computational Investigation of Chemical Reactions ...

60

is observed (R2 = 0.96), consistent with the Bell model of force-activated chemistry.102 Similarly

to TS-1, the activation enthalpies of TS-2 decrease linearly with applied force (R2 = 0.99), but

slope of only about one-half that of TS-1. Based on this difference in slopes, we can extrapolate

that at forces greater than 6.45 nN, bond scission is expected occur via PATH-1 over PATH-2.

4.5 Determining the Effective Tensile Strength of PAA With and Without Radical Attack

In bulk materials, the tensile strength of a material is defined as the maximum tension that a

material can withstand before breaking. However, in looking at a single polymer strand, and

considering the thermochemistry of bond scission, bond scission becomes inevitable as long as it

is exergonic. A more informative measure of mechanical strength in bonds is describing the half-

life of the scission process at a given force and temperature. Thus, we define the effective tensile

strength (E)FT, as the amount of force needed to lower the half-life to small enough time that bond

scission would be expected to occur. Notably, (E)FT is a context-dependent quantity. For instance,

applications that occur at longer timescales would need a lower force for bond scission to occur,

as longer half-lives could be useful for that application.

For the purposes of this work, we are most concerned with timescales and temperatures that are

relevant during ultrasonic depolymerization. Bubble collapse has been previously modeled as

occurring within a 1 μs timescale.137 While acoustic bubbles are known to form local hot spots,138

we will focus on the temperature of the bulk solution, and assume a room-temperature reaction.

Local hot-spots are known to reach up to 2,000K, and it would be expected that both 1 and 2 would

break apart instantaneously under those conditions. Additionally, while the enthalpy of

polymerization of PAA is known to be 18.5 kcal/mol,139 the ceiling temperature of polymerization

has not been directly measured. However, based on high-pressure polymerizations of acrylic acid,

it is estimated that the ceiling temperature is somewhere around 200°C.140

Page 79: Computational Investigation of Chemical Reactions ...

61

Figure 4-11. Derivation of a relationship between enthalpy of activation (ΔH‡) and reaction half-life (t1/2). Where: kb

is Boltzmann’s constant, T is the reaction temperature, h is Planck’s constant, R is the gas constant, ΔH‡ and Δ ‡ are

the enthalpy and entropy of activation, ΔHpolym and Δ polym are the enthalpy and entropy of polymerization, and Tc is

the ceiling temperature of polymerization.

Entry 𝑡1/2

in μs 𝑇 in

°C 𝑇𝑐 in

°C ∆𝐻‡ in

kcal/mol (𝐸)𝐹𝑇 of

1 in nN

(𝐸)𝐹𝑇 of 2 in

nN

𝐿𝐿1𝐿𝐿2

1 1 25 200 21.1 4.7 2.5 1.37

2 0.1 25 200 19.8 4.9 2.9 1.30

3 1 100 200 26.6 4.1 1.1 1.93

4 1 25 300 19.1 4.9 3.0 1.28

5 0.1 100 200 24.9 4.3 1.6 1.63 Table 4-1.The effective tensile strength of 1 and 2 calculated under different assumptions.

Using these 3 assumptions, in conjunction with an equation relating ΔH‡ to a given half-life

(derived in Figure 4-11) and the relationship between force and TS energy in Figure 4-10, the

effective tensile strength of 1 and 2 can be estimated (Table 4-1, entry 1). How the calculated

tensile strength is expected to change by changing these assumptions is shown in entries 2-5. In

general, the estimated values of (E)FT are lower than those predicted by the Morse model. Changes

in the timescale assumption have a logarithmic effect (Entry 2), but changes in temperature have

Page 80: Computational Investigation of Chemical Reactions ...

62

a dramatic effect on the tensile strength of the material (Entries 3, 5). Changes in the assumption

of the ceiling temperature of PAA appear to have a relatively small effect (Entry 4).

Figure 4-12. A contour plot of the effective tensile strength (E)FT of 1 and 2 assuming a Tc value of 200 °C.

Based off of the relationship between tensile strength and limiting length derived by Okkuama and

Hirose,119 the expected ratio of limiting lengths should be equivalent to the square root of their

ratio of tensile strengths. Using the assumptions in Table 4-1, entry 1), it is estimated that

sonication of PAA in the absence of radicals (PATH-1) should result in a limiting length (LL) that

is 1.37 times the length of what would be expected to be observed in the presence of radicals

(PATH-2). This ratio is affected by the assumptions made, and the details on how the ratio of

limiting lengths can change is also shown in Table 4-1.

A contour plot of effective tensile strength at various temperatures and timescales is shown in

Figure 8. This plot allows for extrapolation of the effective tensile strength of 1 and 2 at

temperatures and timescales beyond the assumptions made in Table 4-1. The effective tensile

strength of 2 is expected to weaken to 0 nN at about 200 °C on a reaction timescale of 0.1 μs, and

at temperatures higher than 400 °C, is expected to have no appreciable tensile strength on a

Page 81: Computational Investigation of Chemical Reactions ...

63

timescale of 10 ps. This is in contrast with 1, which still requires some amount of force for scission

to occur within 10 μs at all temperatures calculated.

4.6 Geometric Distortions Due to Applied Force on the Reactant and Transition State

Figure 4-13. Comparison of key geometries, which suggest that PATH-1 initial and TS structures converge towards

one another as forces increase, but PATH-2 structures do not. Explicit water molecules are removed for clarity.

The large difference in activation energy slope between PATH-1 and PATH-2 is surprising, given

that the two reactions occur with similar polymer backbones (Figure 4-10). This phenomenon

might be explained by the geometric distortions imposed upon the reactant and transition states for

the two pathways. Figure 4-13 details the change in geometry from starting structure to transition

state for selected snapshots at 3 and 5 nN of applied force. From examination of PATH-1, it is

Page 82: Computational Investigation of Chemical Reactions ...

64

clear that TS-1 becomes more similar to 1 at 5 nN, compared to at 3 nN (Figure 4-13, A and B).

In particular, this is driven by the fact that at 5 nN, the scissile bond distance of TS-1 is shorter,

and the end-to-end bond length of 1 is longer. This represents a stark difference to the changes in

PATH-2 from 3 nN to 5 nN. Comparing 2 with TS-2 at 3 and 5 nN (Figure 4-13, C and D), the

two structures track each other closely as the applied force changes, due elongation of the reactant

and TS (transition state) at about the same rate.

Figure 4-14. The effects of applied force on the end-to-end lengths in PATH-1 and PATH-2.

Quantitatively tracking the average end-to-end and scissile bond distances in PATH-1 and PATH-

2 confirm the trends seen upon visual inspection of the structures. In the case of PATH-1, with

increasing applied force, the ends of the polymer increasingly elongate (Compound 1, Figure 4-14,

A). However, the end-to-end distance of transition state TS-1 remains relatively stationary,

Page 83: Computational Investigation of Chemical Reactions ...

65

increasing by ~0.1 Å (~1% of length) as the applied force increases (Compound TS-1, Figure 4-14,

A). In essence, at higher force loadings, the end-to-end bond distance of 1 converges to that of TS-

1. The starting and transition state structures of PATH-2 do not exhibit this trend. Rather, both

starting structure 2 and transition state structure TS-2 elongate at a similar rate with increasing

applied force (Figure 4-14, B). At comparable forces, the end-to-end bond lengths of 2 are longer

than 1, while the end-to-end bond lengths of TS-2 eventually converge to similar lengths of TS-1

at ~4 nN.

A better representation of the divergence of the end-to-end bond lengths is to observe how it

changes between the starting structure and transition state at different force values (Figure 4-14,

C). With increasing applied force, the end-to-end distance in PATH-1 decreases at a rate of 0.27

Å per nN of applied force. However, while the end-to-end lengths of 2 and TS-2 consistently

increase with force, in PATH-2 the end-to-end distance only increases at a rate of 0.02 Å per nN

of applied force, essentially keeping the change in polymer length constant between 2 and TS-2.

While in PATH-1 the end-to-end length of 1 converges to that of TS-1 with increasing force, we

see the opposite trend in the change in scissile bond length (Figure 4-15, A). The length of the

scissile bond remains relatively unchanged in 1 with increasing force, but decreases substantially

with increasing force with TS-1. In PATH-2, the scissile bond length in TS-2 decreases at a slower

pace compared to TS-1, while the bond length in 2 is essentially unchanging (Figure 4-15, B). In

both cases, we can interpret this as the transition state of bond scission coming earlier and earlier

in the bond scission event, which is consistent with the notion that under force, the structure of

transition state converging towards starting structure.

Page 84: Computational Investigation of Chemical Reactions ...

66

Figure 4-15. The effects of applied force on the scissile bond distances in PATH-1 and PATH-2.

It is noteworthy, however, that if we look at the change in scissile bond length between the

transition state and starting structure at different force values (Figure 4-15, C), the change in

scissile bond length decreases at a much faster rate in PATH-1 than PATH-2. Just as is seen in

the change in end-to-end lengths (Figure 4-15, C), the geometries of structures in PATH-1 appear

to converge together at higher force values, while in PATH-2, the geometric differences between

TS-2 and 2 are mostly maintained, even at high levels of force.

Under the assumption that the transition state does not appreciably change with increasing amount

of applied force, the change in activation barrier with force for a given reaction is thought to be

caused by the energy of applied force (∆𝑬𝑨𝑭, Figure 4-16), which is simply the energetic

Page 85: Computational Investigation of Chemical Reactions ...

67

contribution stemming from the gradient that force creates.133 In the transition state searching,

force was applied directly to the ends of the polymer, so we measure Δx as the difference in end-

to-end distance between the starting structure and the transition state. Figure 4-16 details the

calculated Δ AF for PATH-1 and PATH-2. In PATH-2, we see a linear relationship between

applied force and Δ AF, with a slope of approximately -5.15 kcal/mol per nN. This approximately

mimics the decrease of 3.97 kcal/mol per nN we see in ΔH‡ with increasing force (Figure 4-10).

However, in the case of PATH-1, Δ AF increases with force, at a rate of about 2.89 kcal/mol per

nN. The ability of ∆𝐸𝐴𝐹 to predict the change in activation energy in PATH-2, and inability to

provide the same prediction in PATH-1, suggests that the application of external force does not

heavily distort TS-2, but dramatically distorts TS-1. This observation is consistent with what is

quantified in Figure 4-14 and Figure 4-15.

Figure 4-16. The energetic contribution from applied force (ΔEAF) observed in PATH-1 and PATH-2. The decrease

in ΔH‡ with force in PATH-2 can be explained though the steady increase in ΔEAF.

According to the TPES model, under application of force, the reactant and TS structures are

expected to become more geometrically similar.86,141 Plots A and B in Figure 4-17 detail the

Page 86: Computational Investigation of Chemical Reactions ...

68

normalized change in scissile bond distance and end-to-end length with force. In both PATH-1

and PATH-2, the scissile bond length of the transition state approaches the scissile bond of the

initial structure. However, in PATH-1, this trend is much more pronounced, and we also see a

similar trend in the end-to-end lengths. The geometric distortion due to force in PATH-1 is so

great that the scissile bond lengths and end-to-end distances appear to be quickly converging to a

single point.

Figure 4-17. Comparing the (relative) end-to-end and scissile bond distances of the starting and TS structures in

PATH-1 and PATH-2. The average end-to-end and scissile bond distances were calculated within each reaction path.

4.7 Rationalizing why PATH-1 is More Responsive to Force Than PATH-2

With the application of force, PATH-1 changes much more rapidly than PATH-2, both in terms

of transition state energy, as well as starting and transition state geometry. As “force

responsiveness” is a consequence of the inverse of the 2nd derivative of the PES,142 the large

differences in response to applied force suggest that the energy profiles of bond dissociation in

PATH-1 and PATH-2 have very different curvatures. Indeed, without radical abstraction, bond

scission in PATH-1 follows a dissociative pathway that does not contain a transition state (Figure

4-18). However, upon abstraction of a radical, bond scission in PATH-2 now follows a pathway

with a transition state (Figure 4-18). A simple analytical model of the two reaction pathways can

Page 87: Computational Investigation of Chemical Reactions ...

69

explain this large difference in force responsiveness. PATH-1 is modeled as a Morse potential,

and PATH-2, as a quartic potential. The choice of these functions reflects the presence or absence

of a transition state that is expected at no force. Each potential energy surface is then modified by

a linear force times distance term, representing the shift of the potential profile with tensile forces.

Parametrization of these energy surfaces was performed to fit the zero-force profiles from

atomistic simulations, see the appendix for more details.

Figure 4-18. Differences in the PES curvature of PATH-1 and PATH-2.

The results of the analytical model are shown in Figure 4-19. The bottom right of Figure 4-19

demonstrates that applying tensile forces has a greater effect on the Morse potential than the quartic

potential. To understand this, the top of Figure 4-19 shows that TS-1 is dragged downward by the

applied force, while TS-2 stays roughly constant. At the same time, 1 and 2 increase in energy at

about the same rate with applied force. The net effect is that applied force reduces the barriers for

TS-1 as well as TS-2, but TS-2’s activation barriers move faster. The relative slope (𝐸𝑆𝑙𝑜𝑝𝑒1

𝐸𝑆𝑙𝑜𝑝𝑒2) of

Page 88: Computational Investigation of Chemical Reactions ...

70

2.1 for this highly simplified model is similar to that of the full atomistic model, with a slope of

2.0.

Figure 4-19. (Top) How the Morse and quartic potentials behave under applied force (Bottom Left) Comparing the

predicted TS energies of PATH-1 and PATH-2 using the example Morse and quartic functions. The ratio of the

change in energy with force (ESlope1 vs ESlope2) is very close to the atomistic model. (Bottom Right) Comparing the

displacement of the location of the TS using the example Morse and quartic functions. The ratio of the change in

displacement with force at the last five data points (DispSlope1 vs DispSlope2) is consistent with the atomistic model.

The simplified model results are also geometrically consistent with the full atomistic model (Figure

4-19, Bottom Left). As the amount of applied force increases, the scissile bond length of TS-1 and

TS-2 move closer and closer to the lengths of respective reactant states 1 and 2. Just as is seen in

the atomistic model, TS-1 changes at a much faster rate compared to TS-2. If we take the slope of

the last 5 datapoints of the TS displacement/force relationship for both TS-1 and TS-2 (𝐷𝑖𝑠𝑝𝑆𝑙𝑜𝑝𝑒1

𝐷𝑖𝑠𝑝𝑆𝑙𝑜𝑝𝑒2),

Page 89: Computational Investigation of Chemical Reactions ...

71

we find that TS-1 is displaced at 3.9 times the rate as TS-2, comparable to the 8-fold difference

that we see in the atomistic model (Figure 4-15, raw numbers are included in the appendix).

4.8 Conclusions

In summary, abstraction of a radical from the polymer backbone is predicted to reduce the tensile

strength of the polymer. Abstraction of a radical decreases the enthalpy of scission from 92.7

kcal/mol to 14.2 kcal/mol. Crudely, we can estimate that this 78.5 kcal/mol decrease lowers the

tensile strength of the polymer from 6.6 nN to 2.6 nN. More accurately, by determining the

transition state of bond scission of 1 and 2, we can derive a linear relationship between applied

force and enthalpy of activation. Using this relationship, we can refine our estimations of the

effective tensile force of 1 and 2 to 4.7 nN and 2.5 nN, respectively, depending on assumptions

about the timescale of bubble collapse, reaction temperature during bond scission, and ceiling

temperature of PAA.

Compounds 1 and 2 are observed to behave quite differently under force. During degradation

through PATH-1, the relative geometries of TS-1 and 1 are significantly perturbed, but in PATH-

2, the corresponding states do not change significantly. This has been measured by observing

marked changes in 1 and TS-1, juxtaposed against the fairly minor changes seen in 2 and TS-2.

By comparing the normalized rate of change of the scissile bond and end-to-end distances to the

normalized change in TS energy in PATH-1, we can assert that the decrease in TS energy is due

to the 1 and TS-1 becoming more geometrically similar with greater force. As the relative

geometries of TS-2 and 2 don’t change nearly as much, we find that ΔEAF provides a good

explanation for the decrease in TS energy seen in PATH-2. The difference in behavior between

the two pathways can be explained as a consequence of the differences in the curvature of the PES

of PATH-1 and PATH-2.

Page 90: Computational Investigation of Chemical Reactions ...

72

These finding are relevant, as the limiting length of a polymer is dependent on the polymer’s tensile

strength.119 As free radicals are capable of lowering the tensile strength of the polymer, this implies

that degradation to shorter chains is possible in the presence of free radicals. The analysis presented

here can be offered as an explanation for observations such as those by Koda where suppression

of radical formation decreases the limiting length of the polymer,106 and suggest that new strategies

in polymer degradation are possible by leveraging the synergy between mechanochemistry and

radical attack.

Page 91: Computational Investigation of Chemical Reactions ...

73

Chapter 5: Conclusions and Final Thoughts

5.1 Research Summary

Despite the rugged, mottled nature of the potential energy surface, only a handful of chemical

states determine the thermodynamics and kinetics of a given chemical reaction. The central

promise of computational chemistry is its ability to both identify and evaluate these key chemical

states. Some of the chapters in this work have mainly been focused on key state identification

(Chapters 2, 3), whereas other chapters have focused on understanding and evaluating key states

(Chapter 4). As demonstrated, both processes provide useful insights for further reaction

development.

In Chapter 2, the mechanism of a nickel-catalyzed three-component coupling reaction of an

aldehyde, alkyne, and enoate is explored. Prior work on the related reductive [3+2] cycloadditions

suggested that two distinct reaction mechanisms were possible, either an aldol-first mechanism

involving a 7-membered metallacycle,50 or a ketene-first mechanism involving a ketene

intermediate.51 As both mechanisms had experimental support for related reactions, the actual

nature of the studied three-component coupling remained ambiguous. Computational studies

provided value, then, because the mechanism of the specific reaction of interest could be directly

interrogated. It was found that the aldol-first and ketene-first pathways were both feasible (Figure

5-1. A summary of Chapter 2. Two mechanisms for the three-component coupling reaction were found to be plausible.

The selectivity for a specific mechanism appears to be affected by the placement of a methyl group at different

positions of the enoate.Figure 5-1), but alkoxide elimination in the ketene-first regime is fast enough

Page 92: Computational Investigation of Chemical Reactions ...

74

to outcompete metallacycle isomerization, meaning that a ketene-first mechanism appears to be

more likely. However, the substitution pattern of the enoate appears to play a role in the choice of

mechanism, as in α-

Page 93: Computational Investigation of Chemical Reactions ...

75

substituted enoates, ketene elimination is much more difficult, due to the steric encumbrance

provided by a group in the α-position. With α-substituted enoates, metallacycle isomerization is

now faster than ketene elimination, which can explain the experimental observation of a linear side

product in only α-substituted enoates, that implicates the existence of a 7-membered metallacycle.

Overall, while the ketene-first pathway was found to be generally favored in the study, it exists in

a precarious balance with an aldol-first mechanism, and as the experience with α-substituted

enoates shows, small changes are capable of changing the preferred mechanism of the reaction.

This work suggests that a similar balance may exist with other reductive [3+2] cycloaddition

reactions, which, as it eventually turned out, includes the activation mechanism of fumarate

catalysts detailed in Chapter 3.

Figure 5-1. A summary of Chapter 2. Two mechanisms for the three-component coupling reaction were found to be

plausible. The selectivity for a specific mechanism appears to be affected by the placement of a methyl group at

different positions of the enoate.

Page 94: Computational Investigation of Chemical Reactions ...

76

Chapter 3 focused on the development of air-stable nickel(0) NHC (N-heterocyclic carbene) pre-

catalysts, and explaining why competent catalysts could be developed with the NHC IMes, but not

with the NHC BAC. Fumarates were chosen as a stabilizing ligand, as they were already known

to be air-stable,83 and it was thought that they could be tuned to become more reactive (Figure 5-2,

A). A series of air-stable fumarate catalysts were developed that could catalyze the reductive

coupling of an alkyne with an aldehyde. While it was originally believed that the activation

mechanism of the fumarate catalysts involved the dissociation of both fumarate ligands,

computational analysis of the binding energies of those catalysts determined that such an activation

mechanism is unlikely (Figure 5-2, B). Even the weakest bound fumarate complex studied had a

binding energy of at least 13.8 kcal/mol, indicating that the fumarate catalyst would be expected

to be over 10 billion times slower than a catalyst formed in situ. The high activity of the fumarate

catalysts, coupled with the knowledge that a fumarate could act as a catalyst poison, suggested that

an activation mechanism where the fumarate is consumed must be active (Figure 5-2, C).

Figure 5-2. (A) Original design strategy to develop an air-stable Ni(0) pre-catalyst. (B) No Correlation between

reaction yield and the free energy of ligand exchange was observed. (C) The absence of such a correlation suggests

that a fumarate consumption reaction must be active.

Page 95: Computational Investigation of Chemical Reactions ...

77

Figure 5-3. The insights gained from Chapter 2 explain why complex 3-8-A is active, but complex 3-10-A is not. The

two complexes undergo activation through different mechanisms. 3-8-A goes through an aldol-first mechanism,

allowing for catalyst activation, but 3-10-A undergoes a ketene-first activation, resulting in a stable ester-ligated

complex.

Leveraging this computational insight, experimentalists were able to isolate the products of the

predicted fumarate consumption mechanism. As the fumarate consumption products were similar

to the products of the three-component coupling reaction detailed in Chapter 2, both the aldol-first

and ketene-first activation mechanisms for IMes and BAC fumarate complexes were investigated.

Like the previous case, mechanism selectivity appeared to be driven by barrier for ketene

elimination (Figure 5-3). For IMes complex 3-8-A, the barrier for ketene elimination is large,

allowing the catalyst to undergo activation through an aldol-first mechanism. However, ketene

elimination is facile for BAC catalyst 3-10-A, suggesting a ketene-first mechanism. This

difference proves to be critical, as the presence of a chelating ester moiety means that the BAC

catalyst can form a highly-stable carbocyclic complex, trapping the catalyst. However, the IMes

catalyst never gets trapped as a highly stabilized species, allowing for the catalyst to activate.

Through the investigating the mechanism of the activation process with computational chemistry,

our understanding of the fumarate catalyst was revolutionized. Applying this insight, preliminary

Page 96: Computational Investigation of Chemical Reactions ...

78

studies in changing the fumarate catalyst to prefer an aldol-first activation path over a ketene-first

activation path provide a direction for the future.

Figure 5-4. A summary of Chapter 4. (A) Radical abstraction imparts a significant change in the enthalpy of bond

scission. (B) The barrier for bond scission in PATH-1 and PATH-2. (C) Differences in the geometries of key states

in PATH-1 and PATH-2. (D) We can explain the differences in behavior between PATH-1 and PATH-2 to be due to

the changes in the curvature of the potential energy surfaces of the two processes.

In Chapter 4, the effect of radical abstraction on the force-enabled bond scission of poly(acrylic

acid) was investigated. Radical abstraction was found to reduce the enthalpy of bond scission by

78.5 kcal/mol, from 92.7 kcal/mol (PATH-1, no radical abstraction) to 14.2 kcal/mol (PATH-2,

with radical abstraction) (Figure 5-4, A). By treating the C-C bond as a Morse oscillator, and

assuming bond scission occurs when no barrier exists for it on the potential energy surface, a crude

estimation of the change in tensile strength suggests that radical attack reduces the tensile strength

of PAA from 6.6 nN to 2.6 nN. Using transition state searching methods recently developed in the

Zimmerman group,96 the transition states of bond scission under force were determined for PATH-

1 and PATH-2. PATH-2 was found to generally have a lower barrier for bond scission, the barrier

Page 97: Computational Investigation of Chemical Reactions ...

79

for bond scission in PATH-2 decreased with increasing applied force compared to PATH-1. Using

the linear relationship of bond scission barrier with applied force, effective tensile strengths in

PATH-1 and PATH-2 were calculated to be 4.7 and 2.5 nN, respectively (Figure 5-4, B).

Additionally, the rationale for how force affects the barrier for bond scission was postulated to be

different in both PATH-1 and PATH-2. In PATH-1, the geometries of the starting structures and

transition states of bond scission were found to change substantially upon application of applied

force, whereas geometric distortions in PATH-2 under force are fairly minimal (Figure 5-4, C).

The large geometric distortions under force, pushing the starting structure and transition state

closer together, can rationalize the decrease in bond dissociation energy in PATH-1. The lack of

distortion in PATH-2, on the other hand, can be explained by the energy of applied force, the

energetic benefit of bond lengthening in the transition state under force. The difference in behavior

between PATH-1 and PATH-2 can also be explained as a consequence of the difference in the

curvature of their respective PESs (Figure 5-4, D). The insights discovered in this chapter suggest

a mechanism of synergy between radical attack and application of tensile force, and may aid in the

development of new depolymerization technology.

5.2 Possible Future Directions for the Material Studied in This Work

Chapters 2 and 3 were mainly focused on the identification of the key states involved in the three-

component coupling reaction, or fumarate catalyst activation, rather than understanding their

nature. Now that metallacycle isomerization and ketene elimination have been identified as key

states, it may be a worthwhile endeavor to try to better understand what factors are capable of

stabilizing and destabilizing them. Tools such as energy decomposition analysis,143 orbital

analysis,144 or distortion-interaction analysis145,146 are all possible means of calculating the impact

of specific features on the selectivity between an aldol-first or a ketene-first reaction. In addition

Page 98: Computational Investigation of Chemical Reactions ...

80

to directly computing the effects of specific descriptors, selectivity-determining features could also

be found through statistical analysis.147

While identifying the features that determine mechanism selectivity for the reactions studied in

Chapters 2 and 3 would be of academic interest, the practical interest of such an inquiry would be

in the development of new pre-catalysts. Smaller NHCs represent an attractive target for pre-

catalyst development, as they have been more unreliable in our hands in in situ procedures. The

preliminary computational work to identify an active BAC fumarate catalyst has already yielded

several candidates of interest, and the relatively minor amount of work required to yield those

results suggests that future attempts with BAC and other NHCs will prove to be similarly facile.

Additionally, the computational analysis described in this work have provided multiple hypothesis

that can be easily proven or disproven through experimental work. For instance, in Chapter 3, the

failure of a BAC fumarate catalyst to activate is hypothesized to be due to its proclivity to become

trapped as an organometallic carbocyclic compound. In theory, this species, or species derived

from it, should be isolable in experiment. The success of such an experiment would provide direct

evidence that complexes such as 3-10-A undergoes a ketene-first activation process, but get

trapped in the process. Such a finding would give credence to the approach of developing fumarate

catalysts that go through an aldol-first activation pathway.

Chapter 4 provides many hypotheses that need to be tested as well. For instance, AFM experiments

to test the effect of radical abstraction on tensile strength would directly confirm the central

assertion of Chapter 4. At the time of writing, such an experiment is being actively researched in

the McNeil group. Additionally, calculations of the rate of bond scission through PATH-1 and

PATH-2 give an estimation on the ratio of limiting length expected during the sonication of

poly(acrylic acid) in the presence and the absence of free radicals. This should be directly

Page 99: Computational Investigation of Chemical Reactions ...

81

measurable, either through the use of radical quenching reagents to inhibit the action of any

radicals generated in solution,106 or through the introduction of a radical generating source such as

Fenton’s reagent.110 As computation provides a quantitative prediction of the expected ratio of

limiting lengths, such an experiment would be a useful barometer for our understanding of the role

of radicals in the sonication process.

5.3 On the Frailty of Human Understanding

In a famous philosophy paper, Edmund Gettier attacked the notion that “justified belief”, i.e. the

existence of supporting evidence for a belief, is sufficient to constitute knowledge.148 Gettier

provides counterexamples, where the justification for an individual’s belief is based on a flawed

premise that makes a successful prediction by coincidence, creating the illusion that the

individual’s belief is knowledgeable. While the original “Gettier problem” consisted of a contrived

example concerning the prediction of employment based on the applicant’s pocket’s contents, real-

world examples of Gettier problems are not uncommon in the research process.

For instance, the initial stages of catalyst design in Chapter 3 played out exactly like a Gettier

problem. The original set of IMes fumarate complexes were designed under the false premise that

catalyst activation occurs through a dissociative mechanism. The fumarates were chosen for their

perceived lack of binding affinity, due to their enhanced steric bulk as compared to the original

methyl substrates used by Cavell.83 When the synthesized catalysts were found to be productive,

the result was originally interpreted as a justification for a dissociative mechanism. But that success

was coincidental. The subsequent computational work detailed in this chapter demonstrated that

in order for the catalyst to activate, rather than simply dissociate, the fumarate needed to be

destroyed. We had originally believed in a wrong mechanism, despite making a successful

Page 100: Computational Investigation of Chemical Reactions ...

82

prediction that provided justification for our initial beliefs. Our initial success was not due to our

knowledge of catalyst activation, but due to random chance.

Even the most carefully justified models can fail to capture the entire picture. In reporting the

nickel-catalyzed reductive [3+2] cycloaddition of enoates and alkynes, Montgomery50 and

Ogoshi51 detail distinctly different mechanisms for the formation of product, and both groups

provided solid experimental evidence for their assertions. In the analysis in Chapter 2, both

mechanisms prove to be viable, and the selectivity between the two mechanisms is easily shifted

by small perturbations. How often does one consider the possibility that two slightly different

substrates can react with the same catalyst to make the same corresponding product, but will go

through completely different mechanisms? How different would our understanding of [3+2]

reductive cycloadditions be, if only one set of mechanistic studies was published, rather than two?

5.4 On the Symbiosis Between Theory and Experiment

What is instructive, then, is how falsely held beliefs can be dispelled with a new perspective. For

instance, in Chapter 3, computational analysis was able to propose a different mechanism of

catalyst activation through the calculation of binding energies. Computational chemists are well

equipped to interrogate specific queries, but are less able to answer more general questions. While

analysis of the binding energies of fumarate clearly demonstrated that some activation mechanism

occurred, which consumed the fumarate, that analysis did not indicate what kind of reaction was

occurring. Despite months of work, computational efforts to identify the mechanism of fumarate

consumption were fruitless, and that question was only resolved once some of the products of

fumarate consumption were ultimately isolated experimentally. General queries, such as

determining the result of a given reaction, are much more easily answered through experiment than

through computation.

Page 101: Computational Investigation of Chemical Reactions ...

83

In short, the strengths and deficiencies of experiment and computation naturally complement each

other, and combined together they can enjoy a symbiotic relationship. Such a relationship is

displayed throughout Chapter 3, as the chapter consists of a series of iterations where the

experimentalists feed insight to theoreticians, and vice versa. It is impossible to talk about the

contributions from one side without mentioning the other. The experimentalists would be lost,

chasing after the wrong mechanism of catalyst activation, absent a computational investigation.

The theoretician would be stuck, unable to process all possible methods of fumarate consumption,

without experimental isolation of the products of fumarate consumption. Computational chemistry

has always needed the support of experimentalists, as they ultimately need to test in the real world

the predictions made in simulation. But, as Chapter 3 demonstrates, the growth of computing

power and the development of better tools have made computational chemistry powerful enough

that perhaps the experimentalists need the theoreticians as well.

Page 102: Computational Investigation of Chemical Reactions ...

84

Appendix

A.1 Computational Details for Chapters 2 and 3

Density functional calculations were performed using Q-Chem 3.1.0.0149 for geometry

optimization and frequency calculations, and ORCA 4.0.0.2150 for single point calculations. All

geometries for intermediates and transition states were optimized using the ωB97X density

functional13 and 6-31G(d) basis set.151,152 Energies were refined by applying the ωB97X-D3

density functional11 with the cc-pVTZ basis153,154 and the SMD implicit solvent model155 with

toluene as the solvent (𝜖 2.4) in Chapter 2, or tetrahydrofuran as the solvent in Chapter 3 (𝜖

7.25). Transition state geometries and minimum energy reaction paths were found using the single-

26 and double-ended24,25 growing string methods. Found transition states were subsequently re-

optimized after the initial search. All energies listed are Gibbs free energies with enthalpy and

entropy corrections at 363 K for Chapter 2, or 298 K for Chapter 3. Entropy corrections were

scaled to 50% to account for the difference in entropy between the gas and solvated phases.156 The

effects of low frequency oscillations were reassigned to 50 cm-1 to prevent the highly anharmonic

vibrations from overly influencing the free energy. All intermediates and transition states were

confirmed to have the appropriate number of imaginary frequencies: one for transition states, and

none for intermediates. All geometry optimizations, frequency calculations, were performed with

an SCF convergence tolerance of 10-6. Single point calculations were performed with an SCF

convergence tolerance of 10-8.

Page 103: Computational Investigation of Chemical Reactions ...

85

A.2 Raw Energies of Structures in Chapter 3

The values used to calculate the relative energies reported for the activation IMes complex 3-8-A

are reported in Table A-1. The values used to calculate the relative energies reported for the

activation BAC complex 3-10-A are reported in Table A-2. The values for the small molecules

used for energy balance in to investigate the pathways of both complexes are given in Table A-3.

The values used to calculate the relative energies of ketene elimination and metallacycle

isomerization for various BAC complexes are given in Table A-4.

Electronic Energy

Zero-Point

Correction Entropy

Complex In hartrees In kcal/mol In cal/mol

IMes-I -3777.07273088773 573.922 280.182

IMes-II -3777.06117723177 573.143 283.499

IMes-TS-II-A -3777.04514633182 572.894 288.034

IMes-III-A -3777.05906179391 574.141 279.708

IMes-TS-III-A -3777.04332245166 570.850 290.026

IMes-TS-II-B -3777.03518980009 572.965 284.492

IMes-III-B -3777.04967194343 571.860 288.283

IMes-IV-A -3777.05529741875 572.734 291.940

IMes-V-A -4221.92540429087 643.799 322.100

IMes-TS-V-A -4221.91122019430 645.019 315.399

IMes-VI-A -4221.93699056449 646.246 318.960

IMes-VI-A-THF -4454.43215241548 725.763 344.347

IMes-TS-VI-A -4631.81573109319 729.845 340.705

IMes-VII-A -4631.84843879111 728.486 353.729

IMes-VIII-A -4631.86712697850 727.700 354.000

IMes-TS-VIII-Z -4631.81904752115 725.879 360.842

IMes-IX-Z -4631.88041138076 731.512 346.623

IMes-TS-VIII-A -4631.83086136393 727.761 348.589

IMes-IX-A -4631.88910809406 728.400 354.700 Table A-1. Energetic values used to calculate relative energies in the activation pathway of IMes catalyst 3-8-A.

Page 104: Computational Investigation of Chemical Reactions ...

86

Electronic Energy

Zero-Point

Correction Entropy

Complex In hartrees In kcal/mol In cal/mol

BAC-I -3550.76343327492 578.960 279.691

BAC-II -3550.75674927357 577.728 280.608

BAC-TS-II-A -3550.74925510619 578.010 280.754

BAC-III-A -3550.75561877717 578.742 281.754

BAC-TS-III-A -3550.72892166746 577.582 284.884

BAC-IV-A -3550.74538535549 579.194 280.742

BAC-TS-II-B -3550.73607756652 576.436 287.907

BAC-III-B -3550.75087524428 576.949 286.110

BAC-TS-III-B -3550.74774549022 576.884 284.392

BAC-IV-B -3550.79413870639 579.320 274.808

BAC-TS-IV-B -3960.65643023093 660.998 314.962

BAC-V-B -3205.14638400882 500.812 246.156

BAC-TS-V-B -3205.09188866902 500.374 249.289

BAC-VI-B -3205.15075011543 501.862 248.227 Table A-2. Energetic values used to calculate relative energies in the activation pathway of BAC catalyst 3-10-A.

Electronic Energy

Zero-Point

Correction Entropy

Structure In hartrees In kcal/mol In cal/mol

4-Fluorobenzaldehyde -444.849789364677 69.976 87.066

Tetrahydrofuran -232.476787942352 78.456 73.154

Trimethylsilane -409.886233904854 81.594 81.022

Trimethyl(o-tolyloxy)silane -755.554325411818 159.348 120.207 Table A-3. Energetic values of small molecules used in the calculations of the activation pathways of both 3-8-A

and 3-10-A.

Page 105: Computational Investigation of Chemical Reactions ...

87

Electronic Energy

Zero-Point

Correction Entropy

Step Fumarate In hartrees In kcal/mol In cal/mol

Ref

eren

ce S

truct

ure

(BA

C-I

)

B -3245.91786662683 545.782 262.207

C -3708.05413664578 653.753 307.495

D -3786.65666774105 691.481 310.894

E -3701.18954413056 584.265 291.184

F -3701.18423696250 584.668 291.281

G -3927.93003945740 600.994 302.797

H -3472.12061691597 541.119 266.985

I -3088.62438412179 470.217 241.505

J -3550.75595586925 577.783 283.021

K -4012.88691971857 689.340 318.855

L -4146.30809886054 551.998 295.784

M -3786.65666774105 691.481 310.894

Ket

ene

Eli

min

atio

n

(BA

C-T

S-I

I-B

)

B -3245.8741121607 543.421 266.582

C -3708.0161805980 652.128 307.922

D -3786.6275131312 690.863 313.821

E -3701.1698962121 584.891 293.175

F -3701.1673810115 584.509 289.833

G -3927.9066423323 600.542 312.494

H -3472.0938845943 539.281 274.493

I -3088.5819107818 468.205 246.462

J -3550.7159490670 577.768 286.908

K -4012.8532280731 687.520 321.935

L -4146.2850054481 550.963 303.283

M -3786.6275131312 690.863 313.821

Met

alla

cycl

e Is

om

eriz

atio

n

(BA

C-T

S-I

II-A

)

B -3245.8774350755 544.866 269.403

C -3708.0145582864 652.361 315.516

D -3786.6276568147 691.517 313.872

E -3701.1539418405 585.213 296.583

F -3701.1496653065 584.945 295.110

G -3927.8960549849 601.138 312.019

H -3472.0850514741 540.094 274.022

I -3088.5839657167 469.529 248.021

J -3550.7187597531 579.036 289.319

K -4012.8534021955 688.316 323.586

L -4146.2741266509 550.520 308.544

M -3786.6276568147 691.517 313.872 Table A-4. Energetic values used in the comparison of ketene elimination with metallacycle isomerization of BAC

complexes using different fumarates.

Page 106: Computational Investigation of Chemical Reactions ...

88

A.3 Methodology Used in Chapter 4

To capture the behavior of PAA, an isotactic tetramer was used as a model system (1) (Figure

A-1). The effects of radical abstraction were modeled using the same tetramer, minus an α-

hydrogen next to a central carboxyl group (2) (Figure A-1). Conformers for 1 were generated using

the Confab tool130 in Open Babel.131 8 unique conformers were identified. Conformers for 2 were

generated by abstracting the α-hydrogen from each of the conformers of 1.

Figure A-1. Model PAA systems used in this study.

As hydrogen bonding is known to affect the heat of polymerization of protic polymers such as

PAA,132 and to ensure that the conformers used in this study are consistent with aqueous solvation,

explicit water solvation was used. For each conformer of 1 and 2, the system was surrounded by a

6 Å water shell containing 260 water molecules. Molecular dynamics (MD) simulations were used

to create a series of snapshots of the tetramer in an aqueous environment. These simulations were

carried out using the TINKER package, version 8.2.157 10 ns of NVT molecular dynamics was

simulated with a 1 fs timestep using the CHARMM 2217,18 forcefield with an Andersen thermostat

and a modified Beeman integrator. Custom parameters were created to model the α-radical in 2

(see Section A.4). Water was treated using the TIP3P model. Electrostatic interactions were treated

using Ewald summation, with a cutoff value of 7 Å. All other nonbonded interactions were treated

using a cutoff of 8 Å. The entire system was simulated in a 23.418 Å by 19.166 Å by 18.613 Å

rectangular cuboid box with periodic boundary conditions. After 2 ns of equilibration, snapshots

were taken every 100 fs. For both model systems (1 and 2), the snapshots from each conformer

Page 107: Computational Investigation of Chemical Reactions ...

89

was combined into a single pool of 640 snapshots. The conformers were then relaxed using the

OPTIMIZE function in TINKER.

The geometry of selected snapshots was then re-optimized using a quantum mechanical/molecular

mechanical method (QM/MM). For the QM/MM optimizations, the QM region was treated using

the B3LYP functional12 with the 6-31G(d)151 basis. Edge water molecules were fixed in place. The

MM region was treated using the same CHARMM 2217,18 forcefield also used in the MD

simulations. The Janus model was used for electronic embedding.16,19 Optimizations were

performed with an SCF convergence tolerance 10-6.

Not all of the collected snapshots from the MD simulations were used. For the QM/MM re-

optimization of 1, 200 of the 640 snapshots were used, taken at regular intervals. For 2, snapshots

were screened for those that had a proper alignment of the middle carboxylic acid with the α-

radical. For a β-scission event to occur, the radical must overlap with the scissile bond to

accommodate formation of a new pi bond.158 This was defined as being within 20° of being anti-

periplanar, or 30° of being syn-periplanar with the middle carboxylic acid (see Figure A-1). After

screening all 640 snapshots of 2, 121 were used for QM/MM optimizations. A flowchart for the

entire process is shown in Figure A-2.

Figure A-2. Flowchart for QM/MM optimization and bond scission.

Page 108: Computational Investigation of Chemical Reactions ...

90

After optimizing the geometry of the snapshots, the scissile bond (shown in red, Figure A-2) was

cleaved by applying a stretching force of 10 nN between the two atoms of the scissile bond using

the EFEI (external force explicitly included) method.133 After bond scission, subsequent re-

optimization yielded two PAA dimers (referred to as dimer+dimer snapshots).

As Figure A-2 details, in the case of no radical activation (1), of the 200 snapshots used, 178 were

successfully optimized using the QM/MM method, 170 were then successfully broken, and 166

snapshots were successfully re-optimized as dimer+dimer snapshots. For the no activation case,

dimer+dimer optimization was performed on the triplet surface to prevent recombination, and then

singlepoint energies on the singlet surface were then calculated. In the case of radical activation

(2), of the 121 snapshots used, 103 were successfully optimized using the QM/MM method, 65

were successfully broken, and 62 of the dimer+dimer snapshots were successfully reoptimized.

The enthalpy of bond scission was determined by comparing the total energy of each dimer+dimer

snapshots to the tetramer snapshot it came from. By averaging the difference in energy for each

set of snapshots, a value for the enthalpy of bond scission could be obtained. In the no activation

case (1), the 166 dimer+dimer snapshots were compared to their original tetramer snapshot. The

same analysis was made for the 62 dimer+dimer snapshots in the radical activation case. For all

optimized tetramers in the no activation and radical activation regimes, partial hessian vibrational

analysis134 was performed on the structures absent any explicit water. These values were averaged

to yield the force constant of the scissile bond (shown in red, Figure A-2).

Using the reactant geometry and a C-C scission bond-dissociation coordinate, the single-ended

growing string method allows for the identification of the transition state (TS) of a reaction,

without prior guessing of the structure of the TS.24 Roessler and Zimmerman described a such a

method that includes force biasing (F-GSM).96 Herein, double-ended F-GSM is used to identify

Page 109: Computational Investigation of Chemical Reactions ...

91

the structure of the TS of bond scission of 1, and single-ended F-GSM is used to identify the

structure of the TS of bond scission of 2. A double-ended method was necessary to investigate 1

to calculate a driving coordinate that was sufficient to overcome the large bond dissociation energy

of the reaction. Force was applied at the ends of the tetramer unit, shown in Figure A-3. Due to the

complexity of the system, a tetrameric system was the largest polymer analogue that could be

produce useful data in a reasonable timeframe. Boulatov has previously shown that any effects of

the length of the model system would have a maximum error of ~4 kcal/mol, but that decreases

sharply with increasing system size.159

Figure A-3. Points where force is applied on compounds 1 and 2. The scissile bonds are shown in red.

During ultrasonic depolymerization, the polymer strand will undergo a coil-stretch transition prior

to scission.91 This transition was found to have a confounding effect on the transition state energies

obtained. To resolve this hysteresis problem, structures were optimized at a high level of force,

and then relaxed to their optimized geometries at the desired force level.96

To optimize structures prior to the transition state searching, snapshots were re-optimized with

force applied using the EFEI method.133 For the scission of 1, 129 snapshots of 1 were optimized

at 6 nN, and then re-optimized at forces ranging from 3 to 5 nN. At these force ranges, the bond

scission is sufficiently exothermic for the strings to have an energy profile consistent with having

a transition state. Their corresponding dimer+dimer snapshots were similarly optimized, except

that the procedure was performed on the triplet surface with the scissile atoms frozen. The

tetramer/dimer+dimer pairs were then used as start and end structures to find the transition state

Page 110: Computational Investigation of Chemical Reactions ...

92

of bond scission using double-ended F-GSM. The resulting transition state energies relative to the

starting structure was then averaged to yield a transition state energy at a given force value. A

flowchart from this process is given in Figure A-4.

Figure A-4. Flowchart for identifying the transition state of bond scission for compound 1.

For the scission of 2, tetramers were extended using a force of 6 nN, and then relaxed to their

optimized geometries at the desired levels of force. This ensured that the polymer strand remained

straight, and avoided including the energy of polymer uncoiling in the transition state energy. The

optimized structures were then used as starting points using single-ended F-GSM. The resulting

transition state energies relative to the starting structure was then averaged to yield a transition

state energy at a given force value. A flowchart from this process is given in Figure A-5.

Figure A-5. Flowchart for identifying the transition state of bond scission for compound 2.

Page 111: Computational Investigation of Chemical Reactions ...

93

In order to effectively interrogate the depolymerization of PAA under aqueous conditions, it is

necessary to find a model that yields suitably accurate energetic information. The polymerization

of acrylic acid under aqueous conditions is known to have an enthalpy of polymerization of 18.5

kcal/mol.139 Given that the heat of polymerization of protic monomers is known to be affected by

the hydrogen-bonding ability of the environment, we explored the use of a QM/MM system with

explicit water molecules that were simulated using molecular mechanics. Using a QM/MM method

with explicit water atoms, the value for the enthalpy of polymerization was found to be 15.9

kcal/mol.

Page 112: Computational Investigation of Chemical Reactions ...

94

A.4 Radical Parameters Used in the Molecular Dynamics Simulations in Chapter 4

The following molecular mechanics parameters were used to simulate the radical carbon in

compound 2 for dynamics simulations performed in Chapter 4. They are designed to be used in

conjunction with the CHARMM 22 parameter file (“charmm22.prm”) included in TINKER,

version 8.2.157

atom 999 99 RAD "Radical" 6 12.011 3

vdw 99 2.0900 -0.0680

charge 999 -0.0900

bond 99 42 300.00 1.4800

bond 99 13 365.00 1.5020

bond 99 14 365.00 1.5020

bond 99 1 36.50 1.1000

bond 99 12 440.00 1.4890

angle 42 99 14 65.00 123.50

angle 99 14 13 32.00 112.20

angle 99 14 1 45.00 111.50

angle 14 99 1 40.00 116.00

angle 99 42 52 75.00 126.00

angle 99 42 35 55.00 110.50

angle 42 99 1 32.00 122.00

angle 12 99 14 65.00 123.50

angle 1 99 12 52.00 119.50

angle 14 99 14 27.00 114.00

torsion 14 99 42 52 1.4000 180.00 2

torsion 14 99 42 35 1.4000 180.00 2

torsion 42 99 14 13 0.3000 0.00 3

torsion 42 99 14 1 0.0000 0.00 3

torsion 99 42 35 3 2.0500 180.000 2

torsion 42 13 14 99 0.2000 0.00 3

torsion 15 13 14 99 0.2000 0.00 3

torsion 14 13 14 99 0.2000 0.00 3

torsion 1 13 14 99 0.2000 0.00 3

torsion 42 99 52 35 53.0000 0.00 0

torsion 14 13 14 99 0.2000 0.00 3

torsion 13 14 99 1 0.0000 0.00 3

torsion 1 99 14 1 0.0000 0.00 3

torsion 1 99 42 52 0.0000 180.00 2

torsion 1 99 42 35 0.0000 180.00 2

torsion 13 14 99 12 0.2000 0.00 3

torsion 12 13 14 99 0.2000 0.00 3

torsion 12 99 14 1 0.0000 180.00 3

torsion 14 99 14 1 0.0000 0.00 3

torsion 13 14 99 14 8.5000 180.0

Page 113: Computational Investigation of Chemical Reactions ...

95

A.5 Data Used to Calculate the Tensile Strength and Rates of Bond Scission in Chapter 4

To calculate the tensile strength using the Morse method, the force constant of the scissile bond,

κ, as well as the energy of bond dissociation, E0 need to be calculated. Averages of those values

are listed in Table A-5 and Table A-6.

Avg. E0 (kcal/mol) Std. Err. Std. Dev. N

PATH-1 92.722 0.6126 7.8925 166

PATH-2 14.199 0.7763 6.1126 62 Table A-5. Average calculated bond dissociation energy values in PATH-1 and PATH-2.

Avg. κ (nN/Å) Std. Err. Std. Dev. N

PATH-1 54.0700 0.076961 1.032538 180

PATH-2 54.0471 0.096577 0.970591 101 Table A-6. Average calculated force constants in PATH-1 and PATH-2.

Table A-7 tabulates the average barriers for bond scission in PATH-1 and PATH-2.

Force in nN Avg. ΔH‡ in kcal/mol Std. Err. Std. Dev. N

PA

TH

-1 3 37.212 2.109 15.644 55

3.5 30.568 2.547 18.013 50

4 28.551 2.101 16.546 62

4.5 21.221 1.476 10.440 50

5 19.629 1.805 13.388 55

PA

TH

-2

0 30.236 0.637 3.824 36

1 27.377 0.676 4.585 46

2 24.200 0.767 5.426 50

3 19.959 0.528 4.288 66

4 15.708 0.699 5.326 58

5 10.301 0.473 3.728 62 Table A-7. The average barriers for bond scission in PATH-1 and PATH-2.

Page 114: Computational Investigation of Chemical Reactions ...

96

A.6 Tabulation of the Raw Geometric Data in Chapter 4

The average end-to-end and scissile bond distances of 1, 2, TS-1, and TS-2 are given in Table

A-8. All other geometric analyses performed in Chapter 4 can be derived from these values.

Distances in Å

End-to-End Scissile Bond Force in nN Mean Std. Dev. Mean Std. Dev. N

1

3 9.7764 0.0342 1.5809 0.0023 50

3.5 9.9637 0.0266 1.5918 0.0025 55

4 10.0942 0.0232 1.5934 0.0023 50

4.5 10.2883 0.0239 1.5977 0.0024 62

5 10.4158 0.0234 1.6079 0.0025 55

TS

-1

3 10.9265 0.0625 2.6332 0.0644 50

3.5 10.9073 0.0622 2.4355 0.0654 55

4 10.8884 0.0526 2.3239 0.0609 50

4.5 11.0082 0.0543 2.2566 0.0625 62

5 11.0116 0.0476 2.1550 0.0591 55

2

0 8.7612 0.1050 1.5608 0.0013 36

1 9.5658 0.0532 1.5675 0.0014 46

2 9.9411 0.0407 1.5748 0.0014 50

3 10.2539 0.0263 1.5828 0.0013 66

4 10.5073 0.0211 1.5918 0.0014 58

5 10.6916 0.0220 1.6018 0.0014 62

TS

-2

0 8.9982 0.1097 2.3756 0.0167 36

1 9.8525 0.0584 2.3008 0.0132 46

2 10.2824 0.0499 2.2657 0.0192 50

3 10.5872 0.0314 2.2585 0.0108 66

4 10.8804 0.0283 2.2614 0.0133 58

5 11.0357 0.0287 2.2306 0.0120 62

Table A-8. The average end-to-end and scissile bond distances in 1, 2, TS-1, and TS-2.

Page 115: Computational Investigation of Chemical Reactions ...

97

A.7 Raw Data and Parameters Used in the Mathematical Model in Chapter 4

The parameters used in the mathematical modelling of PATH-1 and PATH-2 in Chapter 4 are

detailed in Table A-9.

PATH-1: 𝐸 𝐷(1 − 𝑒𝛼(𝑟−𝑟𝑒))2 − 𝐹𝑟 PATH-2: 𝐸 𝐴𝑟4 − 𝐵𝑟2 𝐶 − 𝐹𝑟

Parameter Value Parameter Value

𝐷 93 A 71

𝛼 2.63 B 93

re 0.8 C 30.8 Table A-9. Parameters used for mathematical modeling in Chapter 4.

The datapoints for the change in transition state energy for the mathematical models of PATH-1

and PATH-2 are shown in Table A-10.

Value for ∆𝐻‡

Force

PATH-1

(Morse)

PATH-2

(Quartic)

0 93.000 30.454

1 90.266 29.649

2 88.061 28.852

3 86.056 28.063

4 84.182 27.282

5 82.405 26.509

6 80.707 25.745

7 79.074 24.989

8 77.497 24.241

9 75.969 23.501

10 74.486 22.770

11 73.042 22.047

12 71.634 21.333

13 70.260 20.628

14 68.917 19.931

15 67.603 19.242

16 66.315 18.563

17 65.054 17.892

18 63.816 17.230

19 62.601 16.577

20 61.408 15.933 Table A-10.The change in the transition state energy of the mathematical models of PATH-1 and PATH-2.

Page 116: Computational Investigation of Chemical Reactions ...

98

The datapoints for the displacement of the location of the transition state are shown in Table

A-11.

Displacement of Transition

State From F=1

Force

PATH-1

(Morse)

PATH-2

(Quartic)

1 0.0000 0.0000

2 -0.2643 -0.0054

3 -0.4193 -0.0108

4 -0.5295 -0.0161

5 -0.6151 -0.0215

6 -0.6852 -0.0269

7 -0.7447 -0.0323

8 -0.7963 -0.0378

9 -0.8419 -0.0432

10 -0.8828 -0.0486

11 -0.9198 -0.0541

12 -0.9537 -0.0596

13 -0.9850 -0.0650

14 -1.0140 -0.0706

15 -1.0411 -0.0761

16 -1.0665 -0.0817

17 -1.0904 -0.0872

18 -1.1131 -0.0928

19 -1.1345 -0.0985

20 -1.1549 -0.1042 Table A-11. Displacement of the position of the transition state in the mathematical models of PATH-1 and PATH-

2.

Page 117: Computational Investigation of Chemical Reactions ...

99

Bibliography

(1) Becke, A. D. Perspective: Fifty Years of Density-Functional Theory in Chemical Physics.

J. Chem. Phys. 2014, 140 (18), 18A301. https://doi.org/10.1063/1.4869598.

(2) Dewyer, A. L.; Argüelles, A. J.; Zimmerman, P. M. Methods for Exploring Reaction

Space in Molecular Systems. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2018, 8 (2), e1354.

https://doi.org/10.1002/wcms.1354.

(3) Murdoch, J. R. What Is the Rate-Limiting Step of a Multistep Reaction? J. Chem. Educ.

1981, 58 (1), 32–36. https://doi.org/10.1021/ed058p32.

(4) Kozuch, S.; Shaik, S. How to Conceptualize Catalytic Cycles? The Energetic Span Model.

Acc. Chem. Res. 2011, 44 (2), 101–110. https://doi.org/10.1021/ar1000956.

(5) Amatore, C.; Jutand, A. Mechanistic and Kinetic Studies of Palladium Catalytic Systems.

J. Organomet. Chem. 1999, 576 (1–2), 254–278. https://doi.org/10.1016/S0022-

328X(98)01063-8.

(6) Kozuch, S.; Shaik, S. A Combined Kinetic-Quantum Mechanical Model for Assessment

of Catalytic Cycles: Application to Cross-Coupling and Heck Reactions. J. Am. Chem.

Soc. 2006, 128 (10), 3355–3365. https://doi.org/10.1021/ja0559146.

(7) Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136 (3B), B864.

https://doi.org/10.1103/PhysRev.136.B864.

(8) Kohn, W.; Sham, L. J. Self-Consistent Equations Including Exchange and Correlation

Effects. Phys. Rev. 1965, 140 (4A), A1133. https://doi.org/10.1103/PhysRev.140.A1133.

(9) Sousa, S. F.; Fernandes, P. A.; Ramos, M. J. General Performance of Density Functionals.

J. Phys. Chem. A 2007, 111 (42), 10439–10452. https://doi.org/10.1021/jp0734474.

(10) Goerigk, L.; Hansen, A.; Bauer, C.; Ehrlich, S.; Najibi, A.; Grimme, S. A Look at the

Density Functional Theory Zoo with the Advanced GMTKN55 Database for General

Main Group Thermochemistry, Kinetics and Noncovalent Interactions. Phys. Chem.

Chem. Phys. 2017, 19 (48), 32184–32215. https://doi.org/10.1039/c7cp04913g.

(11) Lin, Y. S.; Li, G. De; Mao, S. P.; Chai, J. Da. Long-Range Corrected Hybrid Density

Functionals with Improved Dispersion Corrections. J. Chem. Theory Comput. 2013, 9 (1),

263–272. https://doi.org/10.1021/ct300715s.

(12) Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J. Ab Initio Calculation of

Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force

Fields. J. Phys. Chem. 1994, 98 (45), 11623–11627.

Page 118: Computational Investigation of Chemical Reactions ...

100

(13) Chai, J.-D.; Head-Gordon, M. Systematic Optimization of Long-Range Corrected Hybrid

Density Functionals. J. Chem. Phys. 2008, 128 (8), 084106.

https://doi.org/10.1063/1.2834918.

(14) Jensen, F. Atomic Orbital Basis Sets. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2013, 3

(3), 273–295. https://doi.org/10.1002/wcms.1123.

(15) Vanommeslaeghe, K.; Guvench, O.; D. MacKerell Jr., A. Molecular Mechanics. Curr.

Pharm. Des. 2014, 20 (20), 3281–3292.

(16) Senn, H. M.; Thiel, W. QM/MM Methods for Biomolecular Systems. Angew. Chemie Int.

Ed. 2009, 48 (7), 1198–1229. https://doi.org/10.1002/anie.200802019.

(17) MacKerell, A. D., J.; Bashford, D.; Bellott, M.; Dunbrack, R. L., J.; Evanseck, J.; Field,

M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; Joseph, D.; Kuchnir, L.; Kuczera, K.; Lau, F.

T. K.; Mattos, C.; Michnick, S.; Ngo, T.; Nguyen, D. T.; Prodhom, B.; Reiher, W. E., I.;

Roux, B.; Schlenkrich, M.; Smith, J.; Stote, R.; Straub, J.; Watanabe, M.; Wiorkiewicz-

Kuczera, J.; Yin, D.; Karplus, M. All-Atom Empirical Potential for Molecular Modeling

and Dynamics Studies of Proteins. J. Phys. Chem. B 1998, 102, 3586–3616.

(18) Foloppe, N.; Alexander D. MacKerell, J. All‐atom Empirical Force Field for Nucleic

Acids: I. Parameter Optimization Based on Small Molecule and Condensed Phase

Macromolecular Target Data. J. Comput. Chem. 2000, 21 (2), 86–104.

https://doi.org/10.1002/(SICI)1096-987X(20000130)21:2<86::AID-JCC2>3.0.CO;2-G.

(19) Shao, Y.; Kong, J. YinYang Atom: A Simple Combined Ab Initio Quantum Mechanical

Molecular Mechanical Model. J. Phys. Chem. A 2007, 111 (18), 3661–3671.

https://doi.org/10.1021/jp067307q.

(20) Cramer, C. J. Essentials of Computational Chemistry : Theories and Models, 2nd ed.;

John Wiley & Sons, West Sussex, England, 2004.

(21) Halgren, T. A.; Lipscomb, W. N. The Synchronous-Transit Method for Determining

Reaction Pathways and Locating Molecular Transition States. Chem. Phys. Lett. 1977, 49

(2), 225–232. https://doi.org/10.1016/0009-2614(77)80574-5.

(22) Mills, G.; Jónsson, H.; Schenter, G. K. Reversible Work Transition State Theory:

Application to Dissociative Adsorption of Hydrogen. Surf. Sci. 1995, 324 (2–3), 305–337.

https://doi.org/10.1016/0039-6028(94)00731-4.

(23) Mills, G.; Jónsson, H. Quantum and Thermal Effects in H2 Dissociative Adsorption:

Evaluation of Free Energy Barriers in Multidimensional Quantum Systems. Phys. Rev.

Lett. 1994, 72 (7), 1124–1127. https://doi.org/10.1103/PhysRevLett.72.1124.

(24) Zimmerman, P. M. Growing String Method with Interpolation and Optimization in

Internal Coordinates: Method and Examples. J. Chem. Phys. 2013, 138 (18), 184102.

https://doi.org/10.1063/1.4804162.

(25) Zimmerman, P. Reliable Transition State Searches Integrated with the Growing String

Method. J. Chem. Theory Comput. 2013, 9 (7), 3043–3050.

Page 119: Computational Investigation of Chemical Reactions ...

101

https://doi.org/10.1021/ct400319w.

(26) Zimmerman, P. M. Single-Ended Transition State Finding with the Growing String

Method. J. Comput. Chem. 2015, 36 (9), 601–611. https://doi.org/10.1002/jcc.23833.

(27) Jenkins, A. D.; Robo, M. T.; Zimmerman, P. M.; Montgomery, J. Nickel-Catalyzed

Three-Component Cycloadditions of Enoates, Alkynes, and Aldehydes. J. Org. Chem.

2020, 85 (5), 2956–2965. https://doi.org/10.1021/acs.joc.9b02446.

(28) Trost, B. M. [3+2] Cycloaddition Approaches to Five-Membered Rings via

Trimethylenemethane and Its Equivalents [New Synthetic Methods (55)]. Angew. Chemie

Int. Ed. English 1986, 25 (1), 1–20. https://doi.org/10.1002/anie.198600013.

(29) Burstein, S. H. The Chemistry, Biology and Pharmacology of the Cyclopentenone

Prostaglandins. Prostaglandins Other Lipid Mediat. 2020, 148, 106408.

https://doi.org/10.1016/j.prostaglandins.2020.106408.

(30) Cao, H.; Chen, F.; Su, C.; Yu, L. Construction of Carbocycles from

Methylenecyclopropanes. Adv. Synth. Catal. 2020, 362 (3), 438–461.

https://doi.org/10.1002/adsc.201900615.

(31) Pattenden, G.; Lee, T. V. Saturated Carbocyclic Ring Synthesis. In General and Synthetic

Methods: Volume 11; Royal Society of Chemistry (RSC), 1989; pp 518–546.

https://doi.org/10.1039/9781847556233-00518.

(32) Pellissier, H. Recent Developments in the Nazarov Process. Tetrahedron 2005, 61 (27),

6479–6517. https://doi.org/10.1016/j.tet.2005.04.014.

(33) Hartley, R. C.; Caldwell, S. T. Novel Methods for the Synthesis of Three-, Four-, Five-,

Six- and Seven-Membered, Saturated and Partially Unsaturated Carbocycles. J. Chem.

Soc. Perkin Trans. 1 2000, No. 4, 477–501. https://doi.org/10.1039/a804421j.

(34) Namyslo, J. C.; Kaufmann, D. E. The Application of Cyclobutane Derivatives in Organic

Synthesis. Chem. Rev. 2003, 103 (4), 1485–1537. https://doi.org/10.1021/cr010010y.

(35) Monfette, S.; Fogg, D. E. Equilibrium Ring-Closing Metathesis. Chem. Rev. 2009, 109

(8), 3783–3816. https://doi.org/10.1021/cr800541y.

(36) Iqbal, J.; Bhatia, B.; Nayyar, N. K. Transition Metal-Promoted Free-Radical Reactions in

Organic Synthesis: The Formation of Carbon-Carbon Bonds. Chem. Rev. 1994, 94 (2),

519–564. https://doi.org/10.1021/cr00026a008.

(37) Gibson (née Thomas), S. E.; Middleton, R. J. The Intramolecular Heck Reaction.

Contemp. Org. Synth. 1996, 3 (6), 447–471. https://doi.org/10.1039/co9960300447.

(38) Reissig, H. U.; Zimmer, R. Donor - Acceptor-Substituted Cyclopropane Derivatives and

Their Application in Organic Synthesis. Chem. Rev. 2003, 103 (4), 1151–1196.

https://doi.org/10.1021/cr010016n.

(39) Masse, C. E.; Panek, J. S. Diastereoselective Reactions of Chiral Allyl- and Allenylsilanes

with Activated C=X π-Bonds. Chem. Rev. 1995, 95 (5), 1293–1316.

https://doi.org/10.1021/cr00037a008.

Page 120: Computational Investigation of Chemical Reactions ...

102

(40) Methot, J. L.; Roush, W. R. Nucleophilic Phosphine Organocatalysis. Adv. Synth. Catal.

2004, 346 (910), 1035–1050. https://doi.org/10.1002/adsc.200404087.

(41) Montgomery, J. Nickel-Catalyzed Reductive Cyclizations and Couplings. Angew. Chemie

Int. Ed. 2004, 43 (30), 3890–3908. https://doi.org/10.1002/anie.200300634.

(42) Suzuki, K.; Urabe, H.; Sato, F. A Novel Tandem Cyclization of 2,7- or 2,8-Bis-

Unsaturated Esters Mediated by (Η2-Propene)TiX2 (X = Cl or O-i-Pr). A Facile

Construction of Bicyclo[3.3.0]Octane, -[4.3.0]Nonane, and -[3.1.0]Hexane Skeletons. J.

Am. Chem. Soc. 1996, 118 (36), 8729–8730. https://doi.org/10.1021/ja961928n.

(43) Urabe, H.; Suzuki, K.; Sato, F. Intramolecular Cyclization of 2,7- or 2,8-Bis-Unsaturated

Esters Mediated by (Η2-Propene)Ti(O-i-Pr)2. Facile Construction of Mono- and Bicyclic

Skeletons with Stereoselective Introduction of a Side Chain. A Synthesis of d-Sabinene. J.

Am. Chem. Soc. 1997, 119 (42), 10014–10027. https://doi.org/10.1021/ja9716160.

(44) Hata, T.; Hirone, N.; Sujaku, S.; Nakano, K.; Urabe, H. Iron-Mediated Intramolecular

Metalative Cyclization of α,β- Unsaturated Esters and Amides. Versatile One-Pot

Preparation of Bicyclic Ketoesters. Org. Lett. 2008, 10 (21), 5031–5033.

https://doi.org/10.1021/ol802170z.

(45) Hata, T.; Sujaku, S.; Hirone, N.; Nakano, K.; Imoto, J.; Imade, H.; Urabe, H. Iron-

Mediated and -Catalyzed Metalative Cyclization of Electron-Withdrawing-Group-

Substituted Alkynes and Alkenes with Grignard Reagents. Chem. - A Eur. J. 2011, 17

(51), 14593–14602. https://doi.org/10.1002/chem.201101273.

(46) Wei, C.-H.; Mannathan, S.; Cheng, C.-H. Regio- and Enantioselective Cobalt-Catalyzed

Reductive [3+2] Cycloaddition Reaction of Alkynes with Cyclic Enones: A Route to

Bicyclic Tertiary Alcohols. Angew. Chemie Int. Ed. 2012, 51 (42), 10592–10595.

https://doi.org/10.1002/anie.201205115.

(47) Montgomery, J. Nickel-Catalyzed Cyclizations, Couplings, and Cycloadditions Involving

Three Reactive Components. Acc. Chem. Res. 2000, 33 (7), 467–473.

https://doi.org/10.1021/ar990095d.

(48) Chowdhury, S. K.; Amarasinghe, K. K. D.; Heeg, M. J.; Montgomery, J. Diverse

Reactivity Manifolds of Alkynyl Enone- and Alkynyl Enal-Derived Nickelacycles

Discovery of Nickel-Promoted [3+2] and [2+1] Cycloadditions [6]. J. Am. Chem. Soc.

2000, 122 (28), 6775–6776. https://doi.org/10.1021/ja0012624.

(49) Amarasinghe, K. K. D.; Chowdhury, S. K.; Heeg, M. J.; Montgomery, J. Structure of an

Η1 Nickel O-Enolate: Mechanistic Implications in Catalytic Enyne Cyclizations.

Organometallics 2001, 20 (3), 370–372. https://doi.org/10.1021/om0010013.

(50) Jenkins, A. D.; Herath, A.; Song, M.; Montgomery, J. Synthesis of Cyclopentenols and

Cyclopentenones via Nickel-Catalyzed Reductive Cycloaddition. J. Am. Chem. Soc. 2011,

133 (36), 14460–14466. https://doi.org/10.1021/ja206722t.

(51) Ohashi, M.; Taniguchi, T.; Ogoshi, S. Nickel-Catalyzed Formation of Cyclopentenone

Derivatives via the Unique Cycloaddition of α,β-Unsaturated Phenyl Esters with Alkynes.

J. Am. Chem. Soc. 2011, 133 (38), 14900–14903. https://doi.org/10.1021/ja2059999.

Page 121: Computational Investigation of Chemical Reactions ...

103

(52) Li, W.; Herath, A.; Montgomery, J. Evolution of Efficient Strategies for Enone-Alkyne

and Enal-Alkyne Reductive Couplings. J. Am. Chem. Soc. 2009, 131 (46), 17024–17029.

https://doi.org/10.1021/ja9083607.

(53) Herath, A.; Thompson, B. B.; Montgomery, J. Catalytic Intermolecular Reductive

Coupling of Enones and Alkynes. J. Am. Chem. Soc. 2007, 129 (28), 8712–8713.

https://doi.org/10.1021/ja073300q.

(54) Jackson, E. P.; Malik, H. A.; Sormunen, G. J.; Baxter, R. D.; Liu, P.; Wang, H.; Shareef,

A. R.; Montgomery, J. Mechanistic Basis for Regioselection and Regiodivergence in

Nickel-Catalyzed Reductive Couplings. Acc. Chem. Res. 2015, 48 (6), 1736–1745.

https://doi.org/10.1021/acs.accounts.5b00096.

(55) Jenkins, A. D. Optimization and Development of Nickel-Catalyzed [3+2]-Reductive and

Alkylative Cycloadditions, Univeristy of Michigan: Ann Arbor, MI, 2013.

(56) Nett, A. J.; Canellas, S.; Higuchi, Y.; Robo, M. T.; Kochkodan, J. M.; Haynes, M. T.;

Kampf, J. W.; Montgomery, J. Stable, Well-Defined Nickel(0) Catalysts for Catalytic C-C

and C-N Bond Formation. ACS Catal. 2018, 8 (7), 6606–6611.

https://doi.org/10.1021/acscatal.8b02187.

(57) Tasker, S. Z.; Standley, E. A.; Jamison, T. F. Recent Advances in Homogeneous Nickel

Catalysis. Nature 2014, 509 (7500), 299–309. https://doi.org/10.1038/nature13274.

(58) Prakasham, A. P.; Ghosh, P. Nickel N-Heterocyclic Carbene Complexes and Their Utility

in Homogeneous Catalysis. Inorganica Chim. Acta 2015, 431, 61–100.

https://doi.org/10.1016/j.ica.2014.11.005.

(59) Arévalo, A.; García, J. J. Bond Activation with Low-Valent Nickel in Homogeneous

Systems. Eur. J. Inorg. Chem. 2010, 2010 (26), 4063–4074.

https://doi.org/10.1002/ejic.201000621.

(60) Iwasaki, T.; Kambe, N. Ni-Catalyzed C-C Couplings Using Alkyl Electrophiles. Top.

Curr. Chem. 2016, 374, 66. https://doi.org/10.1007/s41061-016-0067-6.

(61) Mesganaw, T.; Garg, N. K. Ni- and Fe-Catalyzed Cross-Coupling Reactions of Phenol

Derivatives. Organic Process Research and Development. American Chemical Society

January 18, 2013, pp 29–39. https://doi.org/10.1021/op300236f.

(62) Tollefson, E. J.; Hanna, L. E.; Jarvo, E. R. Stereospecific Nickel-Catalyzed Cross-

Coupling Reactions of Benzylic Ethers and Esters. 2015.

https://doi.org/10.1021/acs.accounts.5b00223.

(63) Jana, R.; Pathak, T. P.; Sigman, M. S. Advances in Transition Metal (Pd,Ni,Fe)-Catalyzed

Cross-Coupling Reactions Using Alkyl-Organometallics as Reaction Partners. Chem. Rev.

2011, 111 (3), 1417–1492. https://doi.org/10.1021/cr100327p.

(64) Lavoie, C. M.; Stradiotto, M. Bisphosphines: A Prominent Ancillary Ligand Class for

Application in Nickel-Catalyzed C-N Cross-Coupling. ACS Catalysis. American Chemical

Society August 3, 2018, pp 7228–7250. https://doi.org/10.1021/acscatal.8b01879.

(65) Ritleng, V.; Henrion, M.; Chetcuti, M. J. Nickel N-Heterocyclic Carbene-Catalyzed C-

Page 122: Computational Investigation of Chemical Reactions ...

104

Heteroatom Bond Formation, Reduction, and Oxidation: Reactions and Mechanistic

Aspects. ACS Catal. 2016, 6 (2), 890–906. https://doi.org/10.1021/acscatal.5b02021.

(66) Tobisu, M.; Chatani, N. Cross-Couplings Using Aryl Ethers via C-O Bond Activation

Enabled by Nickel Catalysts. Acc. Chem. Res. 2015, 48 (6), 1717–1726.

https://doi.org/10.1021/acs.accounts.5b00051.

(67) Gooßen, L. J.; Gooßen, K.; Stanciu, C. C(Aryl) O Activation of Aryl Carboxylates in

Nickel-Catalyzed Biaryl Syntheses. Angew. Chemie Int. Ed. 2009, 48 (20), 3569–3571.

https://doi.org/10.1002/anie.200900329.

(68) Tollefson, E. J.; Hanna, L. E.; Jarvo, E. R. Stereospecific Nickel-Catalyzed Cross-

Coupling Reactions of Benzylic Ethers and Esters. Acc. Chem. Res. 2015, 48 (8), 2344–

2353. https://doi.org/10.1021/acs.accounts.5b00223.

(69) Harry, N. A.; Saranya, S.; Ujwaldev, S. M.; Anilkumar, G. Recent Advances and

Prospects in Nickel-Catalyzed C-H Activation. Catal. Sci. Technol. 2019, 9 (8), 1726–

1743. https://doi.org/10.1039/c9cy00009g.

(70) Pototschnig, G.; Maulide, N.; Schnürch, M. Direct Functionalization of C−H Bonds by

Iron, Nickel, and Cobalt Catalysis. Chem. - A Eur. J. 2017, 23 (39), 9206–9232.

https://doi.org/10.1002/chem.201605657.

(71) Yamaguchi, J.; Muto, K.; Itami, K. Nickel-Catalyzed Aromatic C–H Functionalization.

Top. Curr. Chem. 2016, 374, 4. https://doi.org/10.1007/s41061-016-0053-z.

(72) Guo, L.; Rueping, M. Decarbonylative Cross-Couplings: Nickel Catalyzed Functional

Group Interconversion Strategies for the Construction of Complex Organic Molecules.

Acc. Chem. Res. 2018, 51 (5), 1185–1195. https://doi.org/10.1021/acs.accounts.8b00023.

(73) Standley, E. A.; Tasker, S. Z.; Jensen, K. L.; Jamison, T. F. Nickel Catalysis: Synergy

between Method Development and Total Synthesis. Acc. Chem. Res. 2015, 48 (5), 1503–

1514. https://doi.org/10.1021/acs.accounts.5b00064.

(74) Xu, S.; Kim, E. H.; Wei, A.; Negishi, E. Pd- and Ni-Catalyzed Cross-Coupling Reactions

in the Synthesis of Organic Electronic Materials. Sci. Technol. Adv. Mater. 2014, 15 (4),

44201. https://doi.org/10.1088/1468-6996/15/4/044201.

(75) Johansson Seechurn, C. C. C.; Kitching, M. O.; Colacot, T. J.; Snieckus, V. Palladium-

Catalyzed Cross-Coupling: A Historical Contextual Perspective to the 2010 Nobel Prize.

Angew. Chemie Int. Ed. 2012, 51 (21), 5062–5085.

https://doi.org/10.1002/anie.201107017.

(76) CRC Handbook of Chemistry and Physics, 100th Edition (Internet Version 2019), 100th

Edit.; Rumble, J. R., Ed.; CRC Press/Taylor & Francis: Boca Raton, FL, 2019.

(77) Hazari, N.; Melvin, P. R.; Beromi, M. M. Well-Defined Nickel and Palladium Precatalysts

for Cross-Coupling. Nat. Rev. Chem. 2017, 1 (3), 0025. https://doi.org/10.1038/s41570-

017-0025.

(78) Todd, D. P.; Thompson, B. B.; Nett, A. J.; Montgomery, J. Deoxygenative C-C Bond-

Forming Processes via a Net Four-Electron Reductive Coupling. J. Am. Chem. Soc. 2015,

Page 123: Computational Investigation of Chemical Reactions ...

105

137 (40), 12788–12791. https://doi.org/10.1021/jacs.5b08448.

(79) Macklin, T. K.; Snieckus, V. Directed Ortho Metalation Methodology. The N,N-Dialkyl

Aryl O-Sulfamate as a New Directed Metalation Group and Cross-Coupling Partner for

Grignard Reagents. Org. Lett. 2005, 7 (13), 2519–2522.

https://doi.org/10.1021/ol050393c.

(80) Martin, A. R.; Nelson, D. J.; Meiries, S.; Slawin, A. M. Z.; Nolan, S. P. Efficient C-N and

C-S Bond Formation Using the Highly Active [Ni(Allyl)Cl(IPr* OMe )] Precatalyst.

European J. Org. Chem. 2014, 2014 (15), 3127–3131.

https://doi.org/10.1002/ejoc.201402022.

(81) Wu, J.; Faller, J. W.; Hazari, N.; Schmeier, T. J. Stoichiometric and Catalytic Reactions of

Thermally Stable Nickel(0) NHC Complexes. Organometallics 2012, 31 (3), 806–809.

https://doi.org/10.1021/om300045t.

(82) Iglesias, M. J.; Blandez, J. F.; Fructos, M. R.; Prieto, A.; Álvarez, E.; Belderrain, T. R.;

Nicasio, M. C. Synthesis, Structural Characterization, and Catalytic Activity of

IPrNi(Styrene) 2 in the Amination of Aryl Tosylates. Organometallics 2012, 31 (17),

6312–6316. https://doi.org/10.1021/om300566m.

(83) Clement, N. D.; Cavell, K. J.; Ooi, L. L. Zerovalent N-Heterocyclic Carbene Complexes

of Palladium and Nickel Dimethyl Fumarate: Synthesis, Structure, and Dynamic

Behavior. Organometallics 2006, 25 (17), 4155–4165.

https://doi.org/10.1021/om0602759.

(84) Ho, C.-Y.; Chan, C.-W.; He, L. Catalytic Asymmetric Hydroalkenylation of Vinylarenes:

Electronic Effects of Substrates and Chiral N-Heterocyclic Carbene Ligands. Angew.

Chemie Int. Ed. 2015, 54 (15), 4512–4516. https://doi.org/10.1002/anie.201411882.

(85) Kuchenbeiser, G.; Donnadieu, B.; Bertrand, G. Stable

Bis(Diisopropylamino)Cyclopropenylidene (BAC) as Ligand for Transition Metal

Complexes. J. Organomet. Chem. 2008, 693 (5), 899–904.

https://doi.org/10.1016/j.jorganchem.2007.11.056.

(86) Ribas-Arino, J.; Marx, D. Covalent Mechanochemistry: Theoretical Concepts and

Computational Tools with Applications to Molecular Nanomechanics. Chem. Rev. 2012,

114, 5558–5564. https://doi.org/10.1021/cr200399q.

(87) Blanksby, S. J.; Ellison, G. B. Bond Dissociation Energies of Organic Molecules. Acc.

Chem. Res. 2003, 36 (4), 255–263. https://doi.org/10.1021/ar020230d.

(88) Basedow, A. M.; Ebert, K. H. Ultrasonic Degradation of Polymers in Solution. In Physical

Chemistry; Springer Berlin Heidelberg: Berlin, Heidelberg, 1977; pp 83–148.

(89) Zhurkov, S. N.; Korsukov, V. E. Atomic Mechanism of Fracture of Solid Polymers. J.

Polym. Sci. Polym. Phys. Ed. 1974, 12 (2), 385–398.

https://doi.org/10.1002/pol.1974.180120211.

(90) Mar, B. D.; Qi, H. W.; Liu, F.; Kulik, H. J. Ab Initio Screening Approach for the

Discovery of Lignin Polymer Breaking Pathways. J. Phys. Chem. A 2015, 119 (24), 6551–

6562. https://doi.org/10.1021/acs.jpca.5b03503.

Page 124: Computational Investigation of Chemical Reactions ...

106

(91) Caruso, M. M.; Davis, D. A.; Shen, Q.; Odom, S. A.; Sottos, N. R.; White, S. R.; Moore,

J. S. Mechanically-Induced Chemical Changes in Polymeric Materials. Chem. Rev. 2009,

109 (11), 5733–5798. https://doi.org/10.1021/cr9001353.

(92) Beyer, M. K. The Mechanical Strength of a Covalent Bond Calculated by Density

Functional Theory. J. Chem. Phys. 2000, 112 (17), 7307–7312.

https://doi.org/10.1063/1.481330.

(93) Beyer, M. K. Coupling of Mechanical and Chemical Energy: Proton Affinity as a

Function of External Force. Angew. Chemie Int. Ed. 2003, 42 (40), 4913–4915.

https://doi.org/10.1002/anie.200351748.

(94) Aktah, D.; Frank, I. Breaking Bonds by Mechanical Stress: When Do Electrons Decide for

the Other Side? J. Am. Chem. Soc. 2002, 124 (13), 3402–3406.

https://doi.org/10.1021/ja004010b.

(95) Diesendruck, C. E.; Peterson, G. I.; Kulik, H. J.; Kaitz, J. A.; Mar, B. D.; May, P. A.;

White, S. R.; Martínez, T. J.; Boydston, A. J.; Moore, J. S. Mechanically Triggered

Heterolytic Unzipping of a Low-Ceiling-Temperature Polymer. Nat. Chem. 2014, 6 (7),

623–628. https://doi.org/10.1038/nchem.1938.

(96) Roessler, A. G.; Zimmerman, P. M. Examining the Ways To Bend and Break Reaction

Pathways Using Mechanochemistry. J. Phys. Chem. C 2018, 122, 31.

https://doi.org/10.1021/acs.jpcc.8b00467.

(97) Ellen Macarthur Foundation. The New Plastics Economy. Thinking the Future of Plastics

& Catalysing Action, 2017.

(98) U.S. Environmental Protection Agency, Office of Land and Emergency Management.

Advancing Sustainable Materials Management: 2015 Fact Sheet Assessing Trends in

Material Generation, Recycling, Composting, Combustion with Energy Recovery and

Landfilling in the United States; 2018.

(99) Rahimi, A.; García, J. M. Chemical Recycling of Waste Plastics for New Materials

Production. Nat. Rev. Chem. 2017, 1 (6), 0046. https://doi.org/10.1038/s41570-017-0046.

(100) Flosdorf, E. W.; Chambers, L. A. The Chemical Action of Audible Sound. J. Am. Chem.

Soc. 1933, 55 (7), 3051–3052. https://doi.org/10.1021/ja01334a509.

(101) Gogate, P. R.; Prajapat, A. L. Depolymerization Using Sonochemical Reactors: A Critical

Review. Ultrason. Sonochem. 2015, 27, 480–494.

https://doi.org/10.1016/j.ultsonch.2015.06.019.

(102) Kulik, H. J. MODELING MECHANOCHEMISTRY FROM FIRST PRINCIPLES. In

Reviews in Computational Chemistry, First Edition.; Parrill, A. L., Lipkowitz, K. B., Eds.;

John Wiley & Sons, Inc.: Hoboken, New Jersey, 2018; pp 265–311.

https://doi.org/10.1002/9781119518068.ch6.

(103) Henglein, A. Sonochemistry: Historical Developments and Modern Aspects. Ultrasonics

1987, 25 (1), 6–16. https://doi.org/10.1016/0041-624X(87)90003-5.

(104) Review, T.; Jérôme, F.; Chatel, G.; De Oliveira Vigier, K. Depolymerization of Cellulose

Page 125: Computational Investigation of Chemical Reactions ...

107

to Processable Glucans by Non-Thermal Technologies. Green Chem 2016, 18, 3903.

https://doi.org/10.1039/c6gc00814c.

(105) Holkar, C. R.; Jadhav, A. J.; Pinjari, D. V.; Pandit, A. B. Cavitationally Driven

Transformations: A Technique of Process Intensification. Ind. Eng. Chem. Res. 2019, 58

(15), 5797–5819. https://doi.org/10.1021/acs.iecr.8b04524.

(106) Koda, S.; Taguchi, K.; Futamura, K. Effects of Frequency and a Radical Scavenger on

Ultrasonic Degradation of Water-Soluble Polymers. Ultrason. Sonochem. 2011, 18 (1),

276–281. https://doi.org/10.1016/J.ULTSONCH.2010.06.007.

(107) Petit, A. C.; Noiret, N.; Guezennec, J.; Gondrexon, N.; Colliec-Jouault, S. Ultrasonic

Depolymerization of an Exopolysaccharide Produced by a Bacterium Isolated from a

Deep-Sea Hydrothermal Vent Polychaete Annelid. Ultrason. Sonochem. 2007, 14 (2),

107–112. https://doi.org/10.1016/j.ultsonch.2006.03.010.

(108) Zhi, Z.; Chen, J.; Li, S.; Wang, W.; Huang, R.; Liu, D.; Ding, T.; Linhardt, R. J.; Chen, S.;

Ye, X. Fast Preparation of RG-I Enriched Ultra-Low Molecular Weight Pectin by an

Ultrasound Accelerated Fenton Process. Sci. Rep. 2017, 7 (1), 541.

https://doi.org/10.1038/s41598-017-00572-3.

(109) Li, J.; Li, S.; Zheng, Y.; Zhang, H.; Chen, J.; Yan, L.; Ding, T.; Linhardt, R. J.; Orfila, C.;

Liu, D.; Ye, X.; Chen, S. Fast Preparation of Rhamnogalacturonan I Enriched Low

Molecular Weight Pectic Polysaccharide by Ultrasonically Accelerated Metal-Free Fenton

Reaction. Food Hydrocoll. 2018, 95, 551–561.

https://doi.org/10.1016/j.foodhyd.2018.05.025.

(110) Zhi, Z.; Li, J.; Chen, J.; Li, S.; Cheng, H.; Liu, D.; Ye, X.; Linhardt, R. J.; Chen, S.

Preparation of Low Molecular Weight Heparin Using an Ultrasound-Assisted Fenton-

System. Ultrason. Sonochem. 2019, 52, 184–192.

https://doi.org/10.1016/J.ULTSONCH.2018.11.016.

(111) Li, J.; Li, S.; Wu, L.; Yang, H.; Wei, C.; Ding, T.; Linhardt, R. J.; Zheng, X.; Ye, X.;

Chen, S. Ultrasound-Assisted Fast Preparation of Low Molecular Weight Fucosylated

Chondroitin Sulfate with Antitumor Activity. Carbohydr. Polym. 2019, 209, 82–91.

https://doi.org/10.1016/J.CARBPOL.2018.12.061.

(112) Prajapat, A. L.; Gogate, P. R. Intensification of Depolymerization of Aqueous Guar Gum

Using Hydrodynamic Cavitation. Chem. Eng. Process. Process Intensif. 2015, 93, 1–9.

https://doi.org/10.1016/J.CEP.2015.04.002.

(113) Yue, W.; Yao, P.; Wei, Y.; Mo, H. Synergetic Effect of Ozone and Ultrasonic Radiation

on Degradation of Chitosan. Polym. Degrad. Stab. 2008, 93 (10), 1814–1821.

https://doi.org/10.1016/J.POLYMDEGRADSTAB.2008.07.010.

(114) Prajapat, A. L.; Gogate, P. R. Intensification of Depolymerization of Polyacrylic Acid

Solution Using Different Approaches Based on Ultrasound and Solar Irradiation with

Intensification Studies. Ultrason. Sonochem. 2016, 32, 290–299.

https://doi.org/10.1016/J.ULTSONCH.2016.03.022.

(115) Weissler, A. Depolymerization by Ultrasonic Irradiation: The Role of Cavitation. J. Appl.

Page 126: Computational Investigation of Chemical Reactions ...

108

Phys. 1950, 21 (2), 171–173. https://doi.org/10.1063/1.1699618.

(116) Van der hoff, B. M. E.; Gall, C. E. A Method for Following Changes in Molecular Weight

Distributions of Polymers on Degradation: Development and Comparison with Ultrasonic

Degradation Experiments. J. Macromol. Sci. Part A - Chem. 1977, 11 (9), 1739–1758.

https://doi.org/10.1080/00222337708063088.

(117) Henglein, V. A. Die Reaktion Des α, α-Diphenyl-β-Pikryl-Hydrazyls Mit Langkettigen

Freien Radikalen, Die Beim Ultraschallabbau von Polymethacrylsäuremethylester

Gebildet Werden. Die Makromol. Chemie 1955, 15 (1), 188–210.

https://doi.org/10.1002/macp.1955.020150115.

(118) Tabata, M.; Miyazawa, T.; Kobayashi, O.; Sohma, J. Direct Evidence of Main-Chain

Scissions Induced by Ultrasonic Irradiation of Benzene Solutions of Polymers. Chem.

Phys. Lett. 1980, 73 (1), 178–180. https://doi.org/10.1016/0009-2614(80)85230-4.

(119) Okkuama, M.; Hirose, T. Mechanics of Ultrasonic Degradation of Linear High Polymer

and Ultrasonic Cavitation. J. Appl. Polym. Sci. 1963, 7 (2), 591–602.

https://doi.org/10.1002/app.1963.070070216.

(120) Schaefer, M.; Icli, B.; Weder, C.; Lattuada, M.; M Kilbinger, A. F.; Simon, Y. C. The

Role of Mass and Length in the Sonochemistry of Polymers. 2016.

https://doi.org/10.1021/acs.macromol.5b02362.

(121) Okitsu, K.; Cavalieri, F. Chemical and Physical Effects of Acoustic Bubbles. In

Sonochemical Production of Nanomaterials; Springer International Publishing: Cham,

2018; pp 1–17. https://doi.org/10.1007/978-3-319-96734-9_1.

(122) Riesz, P.; Kondo, T.; Krishna, C. M. Sonochemistry of Volatile and Non-Volatile Solutes

in Aqueous Solutions: E.p.r. and Spin Trapping Studies. Ultrasonics 1990, 28 (5), 295–

303. https://doi.org/10.1016/0041-624X(90)90035-M.

(123) Nam, S. N.; Han, S. K.; Kang, J. W.; Choi, H. Kinetics and Mechanisms of the Sonolytic

Destruction of Non-Volatile Organic Compounds: Investigation of the Sonochemical

Reaction Zone Using Several OH · Monitoring Techniques. Ultrason. Sonochem. 2003,

10 (3), 139–147. https://doi.org/10.1016/S1350-4177(03)00085-3.

(124) Makino, K.; Mossoba, M. M.; Riesz, P. Chemical Effects of Ultrasound on Aqueous

Solutions. Formation of Hydroxyl Radicals and Hydrogen Atoms. J. Phys. Chem. 1983,

87 (8), 1369–1377. https://doi.org/10.1021/j100231a020.

(125) Koda, S.; Kimura, T.; Kondo, T.; Mitome, H. A Standard Method to Calibrate

Sonochemical Efficiency of an Individual Reaction System. Ultrason. Sonochem. 2003,

10 (3), 149–156. https://doi.org/10.1016/S1350-4177(03)00084-1.

(126) Kruus, P.; Lawrie, J. A. G.; O’Neill, M. L. Polymerization and Depolymerization by

Ultrasound. Ultrasonics 1988, 26 (6), 352–355. https://doi.org/10.1016/0041-

624X(88)90035-2.

(127) Encina, M. V.; Lissi, E.; Sarasúa, M.; Gargallo, L.; Radic, D. Ultrasonic Degradation of

Polyvinylpyrrolidone: Effect of Peroxide Linkages. J. Polym. Sci. Polym. Lett. Ed. 1980,

18 (12), 757–760. https://doi.org/10.1002/pol.1980.130181201.

Page 127: Computational Investigation of Chemical Reactions ...

109

(128) Berkowski, K. L.; Potisek, S. L.; Hickenboth, C. R.; Moore, J. S. Ultrasound-Induced

Site-Specific Cleavage of Azo-Functionalized Poly(Ethylene Glycol). Macromolecules

2005, 38 (22), 8975–8978. https://doi.org/10.1021/ma051394n.

(129) Busfield, W. K. Heats and Entropies of Polymerization, Ceiling Temperatures,

Equilibrium Monomer Concentrations, and Polymerizability of Heterocyclic Compounds.

In Polymer Handbook; Branrup, J., Immergut, E. H., Eds.; John Wiley & Sons, Inc.: New

York, New York, 1989; p II/295-II/334.

(130) O’Boyle, N. M.; Vandermeersch, T.; Flynn, C. J.; Maguire, A. R.; Hutchison, G. R.

Confab - Systematic Generation of Diverse Low-Energy Conformers. J. Cheminform.

2011, 3 (1), 8. https://doi.org/10.1186/1758-2946-3-8.

(131) O’Boyle, N. M.; Banck, M.; James, C. A.; Morley, C.; Vandermeersch, T.; Hutchison, G.

R. Open Babel: An Open Chemical Toolbox. J. Cheminform. 2011, 3 (10), 33.

https://doi.org/10.1186/1758-2946-3-33.

(132) Sawada, H. Chapter 2. Heat of Polymerization. J. Macromol. Sci. Part C 1969, 3 (2), 339–

356. https://doi.org/10.1080/15583726908545927.

(133) Ribas-Arino, J.; Shiga, M.; Marx, D. Understanding Covalent Mechanochemistry. Angew.

Chemie Int. Ed. 2009, 48 (23), 4190–4193. https://doi.org/10.1002/anie.200900673.

(134) Besley, N. A.; Metcalf, K. A. Computation of the Amide i Band of Polypeptides and

Proteins Using a Partial Hessian Approach. J. Chem. Phys. 2007, 126 (3), 035101.

https://doi.org/10.1063/1.2426344.

(135) Nguyen, T. Q.; Liang, Q. Z.; Kausch, H. H. Kinetics of Ultrasonic and Transient

Elongational Flow Degradation: A Comparative Study. Polymer (Guildf). 1997, 38, 3783–

3793. https://doi.org/10.1016/S0032-3861(96)00950-0.

(136) Moc, J.; Simmie, J. M. Hydrogen Abstraction from N-Butanol by the Hydroxyl Radical:

High Level Ab Initio Study of the Relative Significance of Various Abstraction Channels

and the Role of Weakly Bound Intermediates. J. Phys. Chem. A 2010, 114 (17), 5558–

5564. https://doi.org/10.1021/jp1009065.

(137) Authier, O.; Ouhabaz, H.; Bedogni, S. Modeling of Sonochemistry in Water in the

Presence of Dissolved Carbon Dioxide. Ultrason. Sonochem. 2018, 45, 17–28.

https://doi.org/10.1016/j.ultsonch.2018.02.044.

(138) Suslick, K. S.; Hammerton, D. A.; Cline, R. E. The Sonochemical Hot Spot. J. Am. Chem.

Soc. 1986, 108 (18), 5641–5642. https://doi.org/10.1021/ja00278a055.

(139) Buchholz, F. L. Preparation Methods of Superabsorbent Polyacrylates. In Superabsorbent

Polymers; Buchholz, F. L., Peppas, N. A., Eds.; 1994; pp 27–38.

https://doi.org/10.1021/bk-1994-0573.ch002.

(140) Kinoshita, K.; Takano, Y.; Ohkouchi, N.; Deguchi, S. Free-Radical Polymerization of

Acrylic Acid under Extreme Reaction Conditions Mimicking Deep-Sea Hydrothermal

Vents. ACS Omega 2017, 2, 2765–2769. https://doi.org/10.1021/acsomega.7b00293.

(141) Bell, G. I. Models for the Specific Adhesion of Cells to Cells. Science (80-. ). 1978, 200

Page 128: Computational Investigation of Chemical Reactions ...

110

(4342), 618–627. https://doi.org/10.1126/science.347575.

(142) Konda, S. S. M.; Brantley, J. N.; Bielawski, C. W.; Makarov, D. E. Chemical Reactions

Modulated by Mechanical Stress: Extended Bell Theory. J. Chem. Phys. 2011, 135 (16),

164103. https://doi.org/10.1063/1.3656367.

(143) Hopffgarten, M. von; Frenking, G. Energy Decomposition Analysis. WIREs Comput. Mol.

Sci. 2012, 2 (1), 43–62. https://doi.org/10.1002/wcms.71.

(144) Knizia, G. Intrinsic Atomic Orbitals: An Unbiased Bridge between Quantum Theory and

Chemical Concepts. J. Chem. Theory Comput. 2013, 9 (11), 4834–4843.

https://doi.org/10.1021/ct400687b.

(145) Ess, D. H.; Houk, K. N. Distortion/Interaction Energy Control of 1,3-Dipolar

Cycloaddition Reactivity. J. Am. Chem. Soc. 2007, 129 (35), 10646–10647.

https://doi.org/10.1021/ja0734086.

(146) Chen, S.; Huang, X.; Meggers, E.; Houk, K. N. Origins of Enantioselectivity in

Asymmetric Radical Additions to Octahedral Chiral-at-Rhodium Enolates: A

Computational Study. J. Am. Chem. Soc. 2017, 139 (49), 17902–17907.

https://doi.org/10.1021/jacs.7b08650.

(147) Sigman, M. S.; Harper, K. C.; Bess, E. N.; Milo, A. The Development of

Multidimensional Analysis Tools for Asymmetric Catalysis and Beyond. Acc. Chem. Res.

2016, 49 (6), 1292–1301. https://doi.org/10.1021/acs.accounts.6b00194.

(148) Gettier, E. L. Is Justified True Belief Knowledge? Analysis 1963, 23 (6), 121–123.

https://doi.org/10.1093/analys/23.6.121.

(149) Shao, Y.; Gan, Z.; Epifanovsky, E.; Gilbert, A. T. B.; Wormit, M.; Kussmann, J.; Lange,

A. W.; Behn, A.; Deng, J.; Feng, X.; Ghosh, D.; Goldey, M.; Horn, P. R.; Jacobson, L. D.;

Kaliman, I.; Khaliullin, R. Z.; Kuś, T.; Landau, A.; Liu, J.; Proynov, E. I.; Rhee, Y. M.;

Richard, R. M.; Rohrdanz, M. A.; Steele, R. P.; Sundstrom, E. J.; Woodcock, H. L.;

Zimmerman, P. M.; Zuev, D.; Albrecht, B.; Alguire, E.; Austin, B.; Beran, G. J. O.;

Bernard, Y. A.; Berquist, E.; Brandhorst, K.; Bravaya, K. B.; Brown, S. T.; Casanova, D.;

Chang, C.-M.; Chen, Y.; Chien, S. H.; Closser, K. D.; Crittenden, D. L.; Diedenhofen, M.;

DiStasio, R. A.; Do, H.; Dutoi, A. D.; Edgar, R. G.; Fatehi, S.; Fusti-Molnar, L.; Ghysels,

A.; Golubeva-Zadorozhnaya, A.; Gomes, J.; Hanson-Heine, M. W. D.; Harbach, P. H. P.;

Hauser, A. W.; Hohenstein, E. G.; Holden, Z. C.; Jagau, T.-C.; Ji, H.; Kaduk, B.;

Khistyaev, K.; Kim, J.; Kim, J.; King, R. A.; Klunzinger, P.; Kosenkov, D.; Kowalczyk,

T.; Krauter, C. M.; Lao, K. U.; Laurent, A. D.; Lawler, K. V.; Levchenko, S. V.; Lin, C.

Y.; Liu, F.; Livshits, E.; Lochan, R. C.; Luenser, A.; Manohar, P.; Manzer, S. F.; Mao, S.-

P.; Mardirossian, N.; Marenich, A. V.; Maurer, S. A.; Mayhall, N. J.; Neuscamman, E.;

Oana, C. M.; Olivares-Amaya, R.; O’Neill, D. P.; Parkhill, J. A.; Perrine, T. M.; Peverati,

R.; Prociuk, A.; Rehn, D. R.; Rosta, E.; Russ, N. J.; Sharada, S. M.; Sharma, S.; Small, D.

W.; Sodt, A.; Stein, T.; Stück, D.; Su, Y.-C.; Thom, A. J. W.; Tsuchimochi, T.;

Vanovschi, V.; Vogt, L.; Vydrov, O.; Wang, T.; Watson, M. A.; Wenzel, J.; White, A.;

Williams, C. F.; Yang, J.; Yeganeh, S.; Yost, S. R.; You, Z.-Q.; Zhang, I. Y.; Zhang, X.;

Zhao, Y.; Brooks, B. R.; Chan, G. K. L.; Chipman, D. M.; Cramer, C. J.; Goddard, W. A.;

Gordon, M. S.; Hehre, W. J.; Klamt, A.; Schaefer, H. F.; Schmidt, M. W.; Sherrill, C. D.;

Page 129: Computational Investigation of Chemical Reactions ...

111

Truhlar, D. G.; Warshel, A.; Xu, X.; Aspuru-Guzik, A.; Baer, R.; Bell, A. T.; Besley, N.

A.; Chai, J.-D.; Dreuw, A.; Dunietz, B. D.; Furlani, T. R.; Gwaltney, S. R.; Hsu, C.-P.;

Jung, Y.; Kong, J.; Lambrecht, D. S.; Liang, W.; Ochsenfeld, C.; Rassolov, V. A.;

Slipchenko, L. V.; Subotnik, J. E.; Van Voorhis, T.; Herbert, J. M.; Krylov, A. I.; Gill, P.

M. W.; Head-Gordon, M. Advances in Molecular Quantum Chemistry Contained in the

Q-Chem 4 Program Package. Mol. Phys. 2015, 113 (2), 184–215.

https://doi.org/10.1080/00268976.2014.952696.

(150) Neese, F. Software Update: The ORCA Program System, Version 4.0. WIREs Comput.

Mol. Sci. 2018, 8 (1). https://doi.org/10.1002/wcms.1327.

(151) Hariharan, P. C.; Pople, J. A. The Influence of Polarization Functions on Molecular

Orbital Hydrogenation Energies. Theor. Chim. Acta 1973, 28 (3), 213–222.

https://doi.org/10.1007/BF00533485.

(152) Rassolov, V. A.; Pople, J. A.; Ratner, M. A.; Windus, T. L. 6-31G* Basis Set for Atoms K

through Zn. J. Chem. Phys. 1998, 109 (4), 1223–1229. https://doi.org/10.1063/1.476673.

(153) Dunning, T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. I. The

Atoms Boron through Neon and Hydrogen. J. Chem. Phys. 1989, 90 (2), 1007–1023.

https://doi.org/10.1063/1.456153.

(154) Balabanov, N. B.; Peterson, K. A. Basis Set Limit Electronic Excitation Energies,

Ionization Potentials, and Electron Affinities for the 3d Transition Metal Atoms: Coupled

Cluster and Multireference Methods. J. Chem. Phys. 2006, 125 (7), 074110.

https://doi.org/10.1063/1.2335444.

(155) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on

Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk

Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113 (18),

6378–6396. https://doi.org/10.1021/jp810292n.

(156) Wertz, D. H. Relationship between the Gas-Phase Entropies of Molecules and Their

Entropies of Solvation in Water and 1-Octanol. J. Am. Chem. Soc. 1980, 102 (16), 5316–

5322. https://doi.org/10.1021/ja00536a033.

(157) Rackers, J. A.; Wang, Z.; Lu, C.; Laury, M. L.; Lagardère, L.; Schnieders, M. J.;

Piquemal, J. P.; Ren, P.; Ponder, J. W. Tinker 8: Software Tools for Molecular Design. J.

Chem. Theory Comput. 2018, 14 (10), 5273–5289.

https://doi.org/10.1021/acs.jctc.8b00529.

(158) Carey, F. A.; Sundberg, R. J. Polar Addition and Elimination Reactions. In Advanced

Organic Chemistry: Part A: Structure and Mechanisms; Springer US: Boston, MA, 2007;

pp 473–577. https://doi.org/10.1007/978-0-387-44899-2_5.

(159) Tian, Y.; Boulatov, R. Comparison of the Predictive Performance of the Bell–Evans,

Taylor-Expansion and Statistical-Mechanics Models of Mechanochemistry. Chem.

Commun. 2013, 49 (39), 4187–4189. https://doi.org/10.1039/C2CC37095F.